mapping the cosmological expansion - arxiv · in the remainder of this section we discuss how the...

49
arXiv:0801.2968v2 [astro-ph] 7 Mar 2008 REVIEW ARTICLE Mapping the cosmological expansion Eric V Linder Berkeley Lab & University of California, Berkeley, CA 94720, USA E-mail: [email protected] Abstract. The ability to map the cosmological expansion has developed enormously, spurred by the turning point one decade ago of the discovery of cosmic acceleration. The standard model of cosmology has shifted from a matter dominated, standard gravity, decelerating expansion to the present search for the origin of acceleration in the cosmic expansion. We present a wide ranging review of the tools, challenges, and physical interpretations. The tools include direct measures of cosmic scales through Type Ia supernova luminosity distances, and angular distance scales of baryon acoustic oscillation and cosmic microwave background density perturbations, as well as indirect probes such as the effect of cosmic expansion on the growth of matter density fluctuations. Accurate mapping of the expansion requires understanding of systematic uncertainties in both the measurements and the theoretical framework, but the result will give important clues to the nature of the physics behind accelerating expansion and to the fate of the universe. Contents 1 Introduction 2 1.1 The dynamic universe ............................ 3 1.2 Geometry and destiny ............................ 4 1.3 Acceleration .................................. 6 1.4 Revolution in physics ............................. 7 2 Effects of accelerated expansion 9 2.1 Acceleration directly? ............................. 9 2.2 Kinematics .................................. 10 2.3 Dynamics ................................... 12 2.4 True acceleration? ............................... 12 2.5 Fate of the universe .............................. 15 3 Distance measures 16 3.1 General cosmological distance properties .................. 16 3.2 Type Ia supernovae .............................. 20 3.3 Cosmic microwave background ........................ 21 3.4 Baryon acoustic oscillations ......................... 22

Upload: others

Post on 20-May-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

arX

iv:0

801.

2968

v2 [

astr

o-ph

] 7

Mar

200

8

REVIEW ARTICLE

Mapping the cosmological expansion

Eric V Linder

Berkeley Lab & University of California, Berkeley, CA 94720, USA

E-mail: [email protected]

Abstract. The ability to map the cosmological expansion has developed enormously,

spurred by the turning point one decade ago of the discovery of cosmic acceleration.

The standard model of cosmology has shifted from a matter dominated, standard

gravity, decelerating expansion to the present search for the origin of acceleration in

the cosmic expansion. We present a wide ranging review of the tools, challenges,

and physical interpretations. The tools include direct measures of cosmic scales

through Type Ia supernova luminosity distances, and angular distance scales of baryon

acoustic oscillation and cosmic microwave background density perturbations, as well as

indirect probes such as the effect of cosmic expansion on the growth of matter density

fluctuations. Accurate mapping of the expansion requires understanding of systematic

uncertainties in both the measurements and the theoretical framework, but the result

will give important clues to the nature of the physics behind accelerating expansion

and to the fate of the universe.

Contents

1 Introduction 2

1.1 The dynamic universe . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.2 Geometry and destiny . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

1.3 Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

1.4 Revolution in physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2 Effects of accelerated expansion 9

2.1 Acceleration directly? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.2 Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.3 Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.4 True acceleration? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.5 Fate of the universe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

3 Distance measures 16

3.1 General cosmological distance properties . . . . . . . . . . . . . . . . . . 16

3.2 Type Ia supernovae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.3 Cosmic microwave background . . . . . . . . . . . . . . . . . . . . . . . . 21

3.4 Baryon acoustic oscillations . . . . . . . . . . . . . . . . . . . . . . . . . 22

CONTENTS 2

3.5 Other methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

3.6 Summary of mapping techniques . . . . . . . . . . . . . . . . . . . . . . 26

4 Growth and expansion 26

4.1 Growth of density perturbations . . . . . . . . . . . . . . . . . . . . . . . 26

4.2 Abundance tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

4.3 Gravitational lensing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

4.4 Testing gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

4.4.1 First steps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

4.4.2 Problems parametrizing beyond-Einstein gravity . . . . . . . . . 31

4.4.3 Levels of discovery . . . . . . . . . . . . . . . . . . . . . . . . . . 32

5 Systematics in data and theory 33

5.1 Parameterizing dark energy . . . . . . . . . . . . . . . . . . . . . . . . . 33

5.2 Mirage of Lambda . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

5.3 Inhomogeneous data sets . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

5.4 Miscalibrated standard . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

5.5 Malmquist bias . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

5.6 Other issues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

6 Future prospects 42

6.1 Data and systematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

6.2 Mapping resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

6.3 Limits on cosmic doomsday . . . . . . . . . . . . . . . . . . . . . . . . . 43

7 Conclusions 45

1. Introduction

A century ago our picture of the cosmos was of a small, young, and static universe.

Today we have a far grander and richer universe to inhabit, one that carries information

on the strongest and weakest forces in nature, whose history runs from singularities

and densities and temperatures far beyond our terrestrial and laboratory access to the

vacuum and temperatures near absolute zero. Understanding our universe relies on

a wide range of physics fields including thermodynamics, classical and quantum field

theory, particle physics, and gravitation.

Perhaps most amazing is that while our universe is huge, it is finite in well defined

ways, and can be encompassed and comprehended. The visible universe extends for

nearly an equal number of orders of magnitude above the size of the Earth as the proton

lies below. The number of particles is of order 1080, large but not unbounded, and the

age of the universe is some 14 billion years, only three times the age of the Earth. And

the universe is simple in ways we have no right to expect: it is essentially electrically

neutral and its large scale dynamics is governed by gravity and no other forces of nature,

CONTENTS 3

it is nearly in thermal equilibrium for most of its history, and the geometry of space is

maximally symmetric.

A few characteristics hold where we might be tempted to say, as Alfonso X “The

Wise” did in the 13th century: “Had I been present at the creation of the world, I should

have recommended something simpler”. The universe has about a billion times more

entropy per baryon than expected, and in a related sense has a greatly unequal ratio of

matter to antimatter. The dynamics is neither kinetic energy dominated nor potential

energy dominated but apparently perfectly balanced, giving the critical energy density

and a flat spatial geometry. But the most important cosmological discovery of the 20th

century was that the maximal symmetry does not extend to spacetime; that is, we do

not live in a steady state universe unchanging in time.

Discovery of the cosmic expansion of space in the 1910s and 1920s and that the

evolution arose from a hot, dense, early state called the Big Bang in the 1960s gave rise

to modern cosmology as a exemplar of, window on, and laboratory for physics. Ten

years ago the discovery of the acceleration of that expansion revolutionized cosmology

and a great array of overlapping fields of physics. This article addresses our current

knowledge of the cosmic expansion and our prospects for exploring it in detail. In

particular, we focus on the recent epoch of acceleration and the astrophysical tools for

mapping the cosmic expansion history to reveal the nature of new physics beyond our

present standard model.

In the remainder of this section we discuss how the expansion of our universe

impacts fundamental questions of the origin and fate of the cosmos and everything in

it, and its intimate relation with the nature of gravity, as a force itself and reflecting on

unification with quantum theory. We present the effects of acceleration in §2, but do

not go into the root causes of it, which are highly speculative at this time; see, e.g., the

focus issue on dark energy in [1]. Techniques for directly mapping the expansion appear

in §3, while §4 briefly mentions indirect effects of acceleration through the growth of

structure in the universe. Obtaining an accurate map is crucial to our understanding,

and §5 focuses on challenges imposed by systematic uncertainties within the theoretical

interpretation and data analysis. For future measurements these may be the key issues

in advancing our knowledge. Future prospects for mapping the cosmic expansion from

the Big Bang to the final fate are outlined in §6, and §7 summarizes and concludes.

1.1. The dynamic universe

In the 1910s the frequency shift of spectral lines from astronomical sources with respect

to laboratory measurements were observed by Slipher and the dynamics of space was

found by Einstein within his theory of general relativity. While Einstein, together

with de Sitter, acted to counteract the expansion of space and retain a static state

through introduction of the cosmological constant, other researchers in the 1920s such

as Friedmann and Lemaıtre calculated the expansion history of a universe containing

matter and components with pressure, and Weyl recognized that a physical expansion

CONTENTS 4

scaling linearly with distance occurred naturally. The observations of Hubble in 1929

established the expanding universe as the basis of cosmology.

From the 1930s to 1980s, astronomical observations of the behavior of sources at

different redshifts, and the increasing corroboration of redshift as directly related to

distance, confirmed the picture of an expanding universe. (For a collection of some

important early papers, see [2].) Figure 1 shows the evolution of our capabilities to

map the cosmic expansion and the subsequent understanding achieved. Comparison of

theoretical calculations and observations with respect to primordial nucleosynthesis and

the cosmic microwave background (CMB) radiation clarified the model as one arising

from a hot, dense, near singular state given the name Big Bang. However, estimates of

the matter density (and even more so other component contributions, save for radiation

measured directly in the CMB) remained somewhat vague, and in fact little different

from Friedmann’s 1922 work. Data could not make a definitive statement as to whether

the matter and energy density amounted to less than, equal to, or possibly greater than

the critical density needed for a spatially flat, asymptotically static expansion.

This all changed dramatically in 1998 with the discovery of acceleration of the

cosmic expansion, by two independent groups mapping the distance-redshift relation

of Type Ia supernovae [3, 4]. The expansion history behavior jumped from being

somewhere between a critical, matter dominated universe and an open, spatial curvature

dominated universe (approaching an empty universe akin to the Minkowski space of

merely special relativity), and instead shifted toward one with similarities to de Sitter

space governed by a cosmological constant. What was so revolutionary was not that

such behavior merely changes the quantitative aspects of the expansion, illustrated in

Figure 1, but differs qualitatively and fundamentally, breaking the relation between

geometry and destiny.

1.2. Geometry and destiny

For a non-accelerating universe described on large scales by a smooth, homogeneous

and isotropic, i.e. Friedmann-Robertson-Walker model, geometry and destiny are

inextricably linked. Whether the spatial curvature is positive, zero (flat), or

negative correlates one-to-one with whether the expansion is closed (bounded), critical

(asymptotically static), or open (unbounded). However, the Einstein field equations

governing the cosmic expansion involve not only the amount of energy density but the

full energy-momentum contribution. So there is a loophole to escape destiny.

The Friedmann equations (the Einstein equations in a homogeneous, isotropic

model) can be written as

a2 + k − 8πGρa2/3 ≡ a2 + Veff(a) = 0, (1)

where a is the expansion scale factor, k is the spatial curvature, and ρ is the total

energy density. (We consider the effective potential Veff below.) From the continuity (or

conservation of energy-momentum) equation, the total energy density ρ will evolve with

CONTENTS 5

Figure 1. Mapping the expansion history has been a major theme of cosmology,

revealing the constituents and nature of our universe. From early models with diverse

properties (top left panel), scientists narrowed in to a region shown by the green

box corresponding to Big Bang models with a long matter dominated epoch. By the

1980s cosmologists believed the universe corresponded to an Einstein-de Sitter universe

with critical matter density, or perhaps some model more toward the empty universe

curve (top right panel). The remarkable discovery in 1998 of accelerated expansion

showed that the correct model must be similar to the solid black curve, partaking of

characteristics of the de Sitter universe with a cosmological constant Λ. The challenge

now is to further improve our cosmic mapping ability to zoom in on the narrow green

triangular region, revealing the nature of the new physics behind acceleration (bottom

panel). (See [41] for the specific models.)

expansion as ρ ∼ a−3(1+w), where w = p/ρ is the pressure to energy density, or equation

of state, ratio (easily generalized for a time varying ratio). So ρa2 ∼ a−(1+3w).

CONTENTS 6

If w > −1/3 then the sum of the k and ρa2 terms ranges from −∞ to k for all values

of a, i.e. the entire expansion history. Thus, whether a2 ever reaches zero and hence

whether there is a maximum value of a or expansion continues without halt is wholly

governed by the value of the spatial curvature k: is it positive or negative. Geometry

controls density.

However, if w < −1/3 then the ρa2 term grows with a and eventually dominates

as the expansion factor grows large. This strongly negative contribution can overcome

a positive k, so this breaks the link between geometry and destiny.

One can get a visual appreciation for this by considering the k and ρa2 terms

together as making up an effective potential energy (see, e.g., [5]), such as standardly

used in physics analysis to determine whether a system with a certain kinetic energy is

bound or not. Here, the analog of kinetic energy is a2, and we want to know whether

the minimum kinetic energy, a2 = 0, hits the potential energy curve for any value of a,

indicating a maximum expansion factor.

Figure 2 for the effective potential Veff vs. a then illustrates the conditions discussed

above. If w > −1/3 then the effective potential ranges from −∞ to k for all values of

a, i.e. the entire expansion history. Thus, whether the effective potential crosses the

minimum kinetic energy value of zero and hence whether there is a maximum value of

a or expansion continues without halt is wholly governed by the value of the spatial

curvature k: geometry controls density.

However, if w < −1/3 then the second term of the effective potential eventually

dominates as the expansion factor grows large. This strongly negative contribution can

overcome a positive k, so one can in fact have an eternal, positive curvature universe.

Conversely, if the energy density itself goes negative (e.g. due to a negative cosmological

constant) then a finite, negative curvature universe is possible. To determine the crucial

quantity, the value of w, we need to take into account the entire energy-momentum not

just the energy density. From the other Friedmann equation,a

a= −4πG

3(ρ+ 3p) = −4πG

3ρ (1 + 3w), (2)

we see that the condition w < −1/3 is precisely the condition for accelerated expansion,

a > 0. When one of the components of the universe has sufficiently negative pressure and

contributes substantial energy density, such so-called dark energy can cause the total

equation of state to drop below −1/3 and cause acceleration. Destiny then becomes

unhinged from geometry, and we must understand the nature of dark energy to predict

the fate of the universe. We discuss this further in §2.5.

1.3. Acceleration

Einstein’s equivalence principle states that acceleration is curvature is gravity. While

space may (or may not) be flat, spacetime curvature is nonzero in an expanding universe.

Rather than viewing gravity as a force acting at a distance, we can view it as the

curvature of spacetime and particles follow paths of least action (geodesics) in this

curved spacetime. An intuitive way of seeing these deep connections is given in Figure 3.

CONTENTS 7

Figure 2. Effective potentials for universes classified according to their geometry

(sign of k) and whether the cosmic expansion accelerates (w < −1/3) or not. Models

crossing the dotted zero line do not expand forever. For w > −1/3 the geometry

determines the destiny: whether models with physical kinetic energy, hence Veff ≤ 0,

can achieve a → ∞, i.e. expand forever. By contrast, for accelerating universes the

destiny is eternal expansion regardless of geometry – unless the energy density ρ can

go negative.

Thus, mapping the expansion history and its acceleration is equivalent to mapping

spacetime or to exploring the gravitational nature of the universe.

1.4. Revolution in physics

Is the discovery and further exploration of cosmic acceleration truly revolutionary? Let

us ask what is our current understanding of the universe through the Standard Model:

baryons, photons, neutrinos, etc. make up roughly 4% of the energy density of the

universe. Understanding 4% is like one letter out of the alphabet, e.g. reading

The nature of the physics of the accelerating universe is the premier mystery

for this generation of physics. The challenge of a new, dominant, strongly

negative pressure component of the universe is closely tied with the foundations

of quantum theory and the nature of the vacuum; it is central to the theory

of gravitation and the nature of spacetime, possibly even the number of

dimensions. The behavior of dark energy governs the fate of the universe.

We may hope to soon detect and eventually characterize dark matter. What will

be our understanding once we have succeeded with this characterization and employed

it to extend the Standard Model to a new theory of high energy physics such as

CONTENTS 8

Figure 3. A photon experiences a frequency shift from acceleration, gravity,

or spacetime curvature. Here the spacetime diagram illustrates a Pound-Rebka

experiment of a photon propagating from one atom to another at height h above

it. A gravitational field leads to a gravitational redshift z = gh/c2, while equivalently

a uniform acceleration g over a time t leads to a velocity v = gt and a Doppler shift

z = v/c = gt/c = gh/c2. Since a frequency change is equivalent to a change of time

intervals, the horizontal, parallel line segments representing the emitted photon pulse

period t0 and received period t′ cannot be equal. This means the photon propagation

lines (wavy diagonal lines), which should each be at 45 in a flat spacetime diagram,

must not in fact be parallel. Thus acceleration, gravitation, and spacetime curvature

are equivalent.

supersymmetry? Then we will comprehend 25% of the energy density of the universe,

like all the vowels (with y) of the alphabet:

The nature of the physics of the accelerating universe is the premier mystery

for this generation of physics. The challenge of a new, dominant, strongly

negative pressure component of the universe is closely tied with the foundations

of quantum theory and the nature of the vacuum; it is central to the theory

of gravitation and the nature of spacetime, possibly even the number of

dimensions. The behavior of dark energy governs the fate of the universe.

It is apparent that for true understanding we will need to know the nature of dark

energy. Only then will we have a complete and comprehensible picture:

The nature of the physics of the accelerating universe is the premier mystery

for this generation of physics. The challenge of a new, dominant, strongly

negative pressure component of the universe is closely tied with the foundations

of quantum theory and the nature of the vacuum; it is central to the theory

of gravitation and the nature of spacetime, possibly even the number of

dimensions. The behavior of dark energy governs the fate of the universe.

Understanding cosmic acceleration undeniably will extend the frontiers of physics

and rewrite the textbooks. It is not excessive to suggest that we are faced with a

revolution in our understanding of nature as profound as the mystery of blackbody

radiation a century ago. Blackbody radiation taught us the existence of photons and

CONTENTS 9

the structure of atoms; dark energy may lead us to the existence of quantum gravity

and the structure of the vacuum.

2. Effects of accelerated expansion

2.1. Acceleration directly?

Given the importance of cosmic acceleration, is there some way to detect it in itself,

rather than through the curves of the expansion history? Recall Figure 3. There we

motivated the Equivalence Principle by showing how acceleration equals gravity. We can

detect the acceleration through the cosmological version of the gravitational redshift.

Photons (or any signal) will have their frequencies shifted by the acceleration just as

they would by a gravitational field. Of course the expansion of space in itself redshifts

photons, but that is analogous to a velocity or Doppler shift; acceleration will add a

second time derivative of the photon frequency, showing up a drift in the redshift z.

The redshift drift was first discussed by [6] and given in general form by [7]. Analysis

of its use as a cosmological probe, including observational challenges and systematics,

appeared in [8, 9]. The result is that

z = H0(1 + z)−H(z) = a−1 [H0 − a], (3)

where H = a/a is the Hubble parameter and H0 is the present expansion rate, the

Hubble constant. Since the scale factor a = 1/(1+z) we clearly see that z vanishes only

in universes with a = constant. That is, redshift drift is a direct signature of acceleration

(or deceleration). However, since the Hubble time H−1 is more than 10 billion years, the

redshift drifts at only 1 part in 1010 per year, beyond present technology to measure.

Even with a 20 year observational program with stability achieved at the one part

in a billion level, the cosmological leverage of such a measurement is unimpressive.

Moreover, as [8, 9, 10] pointed out, just as peculiar velocities interfere with accurate

redshift measurements, so would peculiar accelerations degrade redshift drift. These

can take the form of endpoint effects, i.e. jitter in the observer or source motion due to

realistically inhomogeneous gravitational fields from mass flows, or propagation effects

such as a stochastic gravitational wave background with energy density as small as 10−17

of critical density would generate significant noise.

As for dynamical effects of dark energy within the solar system or astrophysical

systems, the energy density is simply too low. All the dark energy within the entire

solar system constitutes the energy equivalent of three hours of sunlight at 1 AU, ruling

out any direct effect on orbits. For lensing by black holes, say, the relative contribution

to the deflection of light by a cosmological constant Λ is Λr2/(m/r). Detectable lensing

requires a gravitational potential Φ ∼ m/r >∼ 10−6, and Λ ∼ H2 so it would require a

black hole with mass greater than 1013M⊙ before dark energy would contribute even

1% effect – and at that point the static approximation used here breaks down. Thus,

cosmological expansion remains the practical path to mapping acceleration.

CONTENTS 10

2.2. Kinematics

It is attractive to consider how much information one can extract on the expansion

history from minimal assumptions. If one uses only the geometry of spacetime as

an input, i.e. the Robertson-Walker metric, then one can learn a remarkable amount.

For example, the expansion of space and redshift of photons, hence the decrease in

temperature and density as the universe expands, are directly derived from examination

of the metric. Such properties that do not rely on supplementing the geometry with

equations of motion are referred to as kinematics, while those that depend on the field

equations, i.e. the specific theory of gravity, fall under dynamics.

For example, the relation between conformal distance η and the expansion factor a

is kinematic,

η(a⋆) =∫ 1

a⋆

da

a2H, (4)

but ifH is defined in terms of η, i.e.H = a/a = a−2da/dη, then this becomes a tautology.

To actually evaluate the distance-redshift relation requires a model for H , generally

supplied by the equations of motion, that is the Friedmann equations in terms of the

densities and pressures. Before returning to this, let us consider kinematic signatures

of acceleration.

In the diagrams of Figure 1, acceleration would show up as convexity of the curves,

i.e. the second derivative becomes positive. A clearer and more direct way of seeing

acceleration is by transforming the variables plotted to the conformal horizon scale as a

function of expansion factor; see Figure 4. A size of the visible universe can be defined in

terms of the Hubble length, H−1, where the expansion, or e-folding, rate H = a/a. Since

all lengths expand with the scale factor a, we can transform into conformal, or comoving,

coordinates by dividing by a, thus defining the conformal horizon scale (aH)−1.

Comoving wavelengths would appear as horizontal lines in this plot and so

wavelengths enter the horizon, i.e. fall below the horizon history curve, at early times

as the universe expands. For decelerating expansion, this is the whole story, with

more wavemodes being revealed as the expansion continues. However, for accelerating

universes, the slope of the horizon history curve goes negative and modes can leave the

horizon. This condition d(aH)−1/d ln a ∼ −a < 0 and its consequences are precisely the

principle behind inflation: an accelerating epoch in the early universe.

This diagram demonstrates visually the tight connection between the expansion

history (the value along the curve), acceleration (the slope of the curve), and distances

of objects in the universe (the area under the curve). Thus distances provide a clear

and direct method for mapping the expansion history, and will be treated at length in

§3.However, as stated above, to actually evaluate the curve for an expansion history

one needs a model for the expansion. One could adopt an ad hoc model a(t) or

H(z) or, parameterizing the acceleration directly, q(z) where q = −aa/a2 is called the

deceleration parameter, or even the third derivative j = a2...a /a3 called the jerk. One

CONTENTS 11

Figure 4. This conformal history diagram presents a unified picture of important

properties of the cosmic expansion. Curves represent the expansion history of different

cosmological models (here the cosmological constant Λ, an extra dimension braneworld

scenario, and a vacuum metamorphosis or quantum phase transition). The value of a

point on a curve measures the conformal horizon, basically the size of the universe, as

a function of the cosmic expansion factor. The slope of the curve gives the deceleration

parameter q; if it is positive then the universe is decelerating but if the slope is negative

the universe is accelerating. The area under a curve gives the distance measured by

observers from the present (a = 1 or ln a = 0) to some time in the past. The shading

around the Λ curve shows how well next generation experiments should probe the

cosmic expansion.

runs the danger of substituting the physics of the Einstein equations with some other,

implicit dynamics since adopting a form for q(z) is equivalent to some ad hoc equation

of motion. Explicitly, if one defines H2 = f(ρ), some function of the total density, say,

then the continuity equation leads to q = −1 + (3/2)(1 + w) d ln f/d ln ρ, so choosing

some functional form q(z), or j(z), is choosing an equation of motion. That is, one

has not truly achieved kinematic constraints, only substituted some other unspecified

physics for general relativity.

One method for getting around this is by putting no physics into the form by taking

a Taylor expansion, e.g. q(z) = q0 + q1z [11, 12]. This has limited validity, that is it can

only be applied to mapping the expansion at very low redshifts z ≪ 1 and it is not clear

what has actually been gained. Another method is to allow the data to determine the

form, and the physics, through principal component analysis as in [13, 12]. This has a

broader range of physical validity, but has substantial sensitivity to noise in the data,

since it seeks to extract information on a third derivative in the case of jerk.

CONTENTS 12

2.3. Dynamics

Einstein’s equations provide the dynamics in terms of the energy-momentum

components in the universe. Note that alterations to the form of the gravitational action

also define the dynamics. Unless otherwise specified we consider Friedmann-Robertson-

Walker cosmologies. In this case, the simplest ingredients are the energy density

and pressure of each component, with the components assumed to be noninteracting.

This can be phrased alternately in terms of the present dimensionless energy density

Ωw = 8πGρw(z = 0)/(3H20) and the pressure to density, or equation of state, ratio

w = pw/ρw.

We have seen in §1.2 that the equation of state is central to the relation between

geometry and destiny, and it plays a key role in the dynamics of the expansion history

as well. In Figure 5 we see the very different behaviors for the dark energy density

for different classes of equations of state. The cosmological constant has an unchanging

energy density, and so it defines a unique scale, tiny in comparison to the Planck energy,

and a particular time in cosmic history when it is comparable to the matter density.

These fine tunings give rise to the cosmological constant problem [14, 15, 16]. Dynamical

scalar fields, called quintessence, change their energy density and this may or may not

alleviate the large discrepancy with the Planck scale. One of the two main classes of such

fields [17], thawing fields that evolve away from early cosmological constant behavior,

has energy density that changes little. The other class, freezing fields that evolve toward

late time cosmological constant behavior, can have much greater energy density at early

times. For example, tracking fields may contribute an appreciable fraction of the energy

density over an extended period in the past and may approach the Planck density

at early times, while scaling fields mimic the behavior of the dominant energy density

component, contributing so much to the dynamics that they can be strongly constrained.

Although there is a diversity of dynamical behaviors [18], these fall into distinct

classes of the physics behind acceleration [17]. We see that to reveal the origin of cosmic

acceleration we will require precision mapping of the expansion history over the last

e-fold of expansion, but we may also need to measure the expansion history at early

times. And to predict the future expansion history and the fate of the universe requires

truly understanding the nature of the new physics – for example knowing whether dark

energy will eventually fade away, restoring the link between geometry and destiny.

2.4. True acceleration?

Before proceeding further we can ask whether the Robertson-Walker model of a

homogeneous, isotropic universe is indeed the proper framework for analyzing the

expansion history. Certainly the global dynamics of the expansion follows that

of a Friedmann-Robertson-Walker model (FRW), as the successes of Big Bang

nucleosynthesis, cosmic microwave background radiation measurements, source counts

etc. show [19, 20], but one could imagine smaller scale inhomogeneities affecting the

light propagation by which we measure the expansion history. This has long been

CONTENTS 13

Figure 5. In the recent expansion history, matter and dark energy have contributed

similar amounts of energy density (shown here normalized to the present matter

density) but this is apparently coincidental. Among quintessence models, the energy

density of the cosmological constant or a thawing field differs substantially at early

times from freezing fields such as trackers or scalers, one example of classification of

dark energy models.

known [21, 22, 23] and for stochastic inhomogeneities shown to be unimportant in the

slow motion, weak field limit [24].

To grasp this intuitively, consider that the expansion rate of space is not a single

number but a 3 × 3 matrix over the spatial coordinates. The analog of the Hubble

parameter is (one third) the trace of this matrix, so inhomogeneities capable of altering

the global expansion so as to mimic acceleration generically lead to changes in the other

matrix components. This induces shear or rotation of the same order as the change in

expansion, leading to an appreciably anisotropic universe. Observations however limit

shear and rotation of the expansion to be less than 5 × 10−5–10−6 times the Hubble

term [25].

Thus, to create the illusion of acceleration one would have to carefully arrange the

material contents of the universe, adjusting the density along the line of sight (spherical

symmetry is not wholly necessary with current data quality). Again, this has long been

discussed [26, 27, 28] and indeed changes the distance-redshift relation. The simple

model of [29] poses the problem in its most basic terms, clearly demonstrating its

meaning. It considers an inhomogeneous, matter only universe with a void (α = 0 for

the Dyer-Roeder smoothness parameter) somewhere along the line of sight, extending

from z1 to z1 + ∆z, and finds that the distances to sources lying at higher redshift do

not agree with the FRW relation. Even for very high redshift sources where the cosmic

CONTENTS 14

volume is essentially wholly described by FRW there maintains an asymptotic fractional

distance deviation of ∆z/(1+ z1) (if the void surrounds the observer then the deviation

is of order ∆z2).

Suppose we were to define w(z) in terms of derivatives of distance with respect

to redshift. Then we would find for certain choices of void size and location that

w(z) < −1/3; see Figure 6. Apart from the fact that this would require enormous

voids, it would be a mistake to interpret this as apparent acceleration within this very

basic, matter only model. The analysis is physically inconsistent because it treats the

expansion, e.g.H(z), in two different ways: as dynamics, for example in the friction term

in the Raychaudhuri or beam equation, and as kinematics, through the correspondence

to the differential of the distance, dr/dz ∼ H . We emphasize this point: in models with

inhomogeneities, it is inconsistent to treat dynamics and kinematics the same. See [30]

for further discussion of the proper physical interpretation of acceleration.

Figure 6. Huge voids can give the illusion of an effective negative equation of state

and acceleration, but not the dynamical reality. The solid curve gives the effective

total equation of state if we observe from the center of a void extending out to z = 0.5,

yet interpret data within a smooth FRW model. The dashed curve corresponds to a

void in a shell from z = 0.55− 0.75.

While the previous approach can deliver a mirage of, but not physical, acceleration,

one cannot even obtain such an illusion if one requires other basic aspects of FRW to

hold. From the Raychaudhuri, or beam, equation (see [21]) in FRW generalized to

arbitrary components and smoothness [31], one finds the following two conditions,

Ωm(z) = 1 + 3wtot(z) ; α(z) =1 + wtot

1 + 3wtot

, (5)

CONTENTS 15

must hold for a pure matter, inhomogeneous model to mimic a smooth model with

matter plus an extra component with equation of state w (so the total equation of

state is wtot). These follow from matching the Raychaudhuri equation term by term

between the models, so that the distance-redshift relations will agree. However, the

requirement of positive energy density then immediately demonstrates that an effective

wtot < −1/3 cannot be attained, i.e. acceleration is not possible. Moreover, the matter in

the inhomogeneous dust model cannot consistently obey the usual continuity equation.

Thus, a pure dust model with inhomogeneities described by spatial under- and

overdensities, i.e. α(z), cannot give distances matching an accelerating FRW universe

distance-redshift relation, nor any FRW model without introducing nonstandard

couplings in the matter sector.

Finally, to obtain even the mirage of a perfectly isotropic distance-redshift relation

requires the inhomogeneities to be arranged in a spherically symmetric manner around

the observer, raising issues of our preferred location. If the inhomogeneities are arranged

only stochastically, i.e. do not have an infinite coherence length caused by special

placement, then the effects along the line of sight will average out and the distance-

redshift relation will reflect the true global expansion. Thus, save for hand fashioning a

universe to deceive, observed acceleration is real acceleration.

2.5. Fate of the universe

Given accelerated expansion at present, we still cannot absolutely predict the future

expansion and the fate of the universe. If the acceleration from dark energy continues

then the universe becomes a truly dark, cold, empty place. The light horizon, within

which we can receive signals, grows linearly with time, as always, but the particle

horizon giving the (at some time) causally influenced region grows more quickly in an

accelerating universe (exponentially in the cosmological constant case). Thus, though

formally our observational reach out into the universe increases, we see an ever tinier

fraction of the causal universe. (See [32] for more on horizons.)

Also, although the light horizon expands, in a real sense the visible universe

does “close in” around us not through objects going beyond the horizon but through

their fading away to our sight. For example, in a cosmological constant dominated

universe the redshift of an object at constant comoving distance increases exponentially

with time, so its received flux decreases exponentially and its surface brightness fades

exponentially from the usual (1 + z)4 law. Conversely, the comoving distance to a

fixed redshift decreases exponentially with time, so vanishingly few objects lie within

the volume to any finite redshift and hence have non-infinitesimal flux and surface

brightness. (See [33, 34, 5] for some specific calculations.) Thus our view of the universe

does not so much shrink as darken. (So dark energy is well named.)

The existence of a horizon arising from acceleration, or Rindler horizon [35], brings

another set of physics puzzles, best known for the borderline case of w = −1. In de Sitter

space the horizon causes loss of unitarity and makes it problematic to define particle

CONTENTS 16

interaction probabilities through an S matrix [14]. On the other hand, the horizon

may be instrumental in obviating the Big Rip fate. For a universe with w < −1 the

increasing (conformal) acceleration overcomes all other binding forces, ripping planets,

atoms, etc. apart [36]. However, Unruh radiation, or thermal particle creation from the

horizon, with temperature T ∼ g and hence energy density ρ ∼ T 4 where g = a/a is

the conformal acceleration (so the rip condition is not...a> 0 but (a/a)˙ > 0), should

quickly overwhelm the dark energy (whose density increases as g, not g4). Thus the

horizon should act to either decelerate the universe or bring the expansion to some

non-superaccelerating state [37].

Other possibilities for the fate of the universe include collapse (a → 0) if the dark

energy attains negative values of its potential, as in the linear potential model [38] (also

see §6.3), collapse and bounce into new expansion as in the cyclic model [39], or eternal

but decelerating expansion if the dark energy fades away.

Thus we have seen throughout this section that to distinguish the origin of cosmic

acceleration we may need not only to map accurately the recent expansion history, but

distinguish models of dark energy through their early time behavior and understand

their nature well enough to predict their future evolution. To truly understand our

universe we must map the cosmic expansion from a = 0 to a = ∞ – and possibly back

to a = 0 again.

3. Distance measures

Distance measurements as a function of scale factor or redshift directly map the

expansion history. Here we consider several approaches to these measurements, starting

with the geometric or mostly geometric methods that give the cleanest probes of

the expansion. Many other techniques involving distances exist, also containing

noncosmological quantities. One can categorize probes into ones depending (almost)

exclusively on geometry, ones requiring some knowledge of the mass of objects to

separate out the distance dependence, and ones requiring not only knowledge of mass

but the thermodynamic or hydrodynamical state of the material, i.e. mass+gas probes.

The discussion here focuses on the geometric distance techniques actually implemented

to date, though we do briefly mention some other approaches.

3.1. General cosmological distance properties

Because distances are integrals over the expansion history, which in turn involves an

integral over the equation of state, degeneracies exist between cosmological parameters

contributing to the distance. These are common to all distance measurements.

Figures 7-8 illustrate the sensitivity and degeneracy of various expansion history

measurements to the cosmological parameters. Here d is the luminosity or angular

distance (we consider fractional precisions so the distinguishing factor of (1+z)2 cancels

out), H is the Hubble parameter, or expansion rate, and the quantities d = d (Ωmh2)1/2,

CONTENTS 17

H = H/(Ωmh2)1/2, where h is the dimensionless Hubble constant, are relative to high

redshift.

Figure 7. Fisher sensitivities for the magnitude (5 log d) are plotted vs. redshift for the

case leaving spatial curvature free (left panel) and fixing it to zero, i.e. a flat universe

(right panel). Since only the shapes of the curves matter for degeneracy, parameters

are normalized to make this more evident. Observations out to z ≥ 1.5 are required

to break the degeneracies.

Figure 8. Fisher sensitivities for the Hubble parameter H (left panel) and reduced

Hubble parameter H = H/(Ωmh2)1/2, i.e. Hubble parameter relative to high redshift

(right panel), are plotted vs. redshift, leaving spatial curvature free. The behavior is

similar to that in Figure 7 but the degeneracies are somewhat more severe.

CONTENTS 18

First we note that for a given fractional measurement precision, distances contain

about as much information as the Hubble parameter, despite the extra integral (the

greater lever arm acts to compensate for the integral). That is (assuming flatness for

simplicity),

d =∫

dz/H, (6)

1

d

∂d

∂θ= −

dz

H2

∂H

∂θ

/

dz

H= −

∂ lnH

∂θ

, (7)

for a parameter θ, where angle brackets denote the weighted average. However, surveys

to measure distances to a given precision can be much less time consuming than those

to measure the Hubble parameter.

Concerning the physical interpretation of the partial derivatives, sometimes known

as Fisher sensitivities since they also enter the Fisher information matrix [40], recall

that

δθ = δ(ln d)

/(

∂ ln d

∂θ

)

. (8)

If for some parameter the Fisher partial derivative at some redshift is 0.88, say, then

a 1% measurement gives an unmarginalized uncertainty on the parameter of 0.011. If

the parameter is wa/10.8, say, this means the unmarginalized uncertainty on wa is 0.12.

The higher the denominator in the parameter label, the less sensitivity exists to that

parameter.

The greatest sensitivity is to the energy densities in matter and dark energy; note

that we do not here assume a flat universe. The nature of the dark energy is here

described through the equation of state parametrization w(a) = w0 + wa(1− a), where

w0 is the present value and wa provides a measure of its time variation (also see §5.1).This has been shown to be an excellent approximation to a wide variety of origins for

the acceleration [41]. Sensitivity to wa is quite modest so highly accurate measurements

are required to discover the physics behind acceleration.

Note that degeneracies between parameters, e.g. Ωm and wa, and Ωw and w0, are

strong at redshifts z <∼ 1. Not until z > 1.5 do distance measurements make an

appreciable distinction between these variables. This determines the need for mapping

the expansion history over approximately the last e-fold of expansion, i.e. to a = 1/e or

z = 1.7. Since the flux of distant sources gets redshifted, this requires many observations

to move into the near infrared, which can only be achieved to great accuracy from space.

From the right panel of Figure 7 we see that the degeneracy between the matter

density Ωm and dark energy equation of state time variation wa holds regardless of

whether space is assumed flat or not. The dark energy density Ωw and the dark energy

present equation of state w0 stay highly degenerate in the curvature free case, whether

considering d or d (not shown in the figure).

From the relation between the derivatives of the distance with respect to

cosmological parameters, as a function of redshift, one can intuit the orientation of

confidence contours in various planes. For example, the sensitivity for the matter density

CONTENTS 19

and dark energy density enter with opposite signs, so increasing one can be compensated

by increasing the other, explaining the low densities to high densities diagonal

orientation in the Ωm-ΩΛ plane. Since the sensitivity to Ωm trends monotonically with

redshift, while that of Ωw has a different shape, this implies that observations over a

range of redshifts will rotate the confidence contours, breaking the degeneracies and

tightening the constraints beyond the mere power of added statistics. Similarly one can

see that contours in the Ωm-w0 plane will have negative slope, as will those in the w0-waplane.

For the tilde variable d, that is distances relative to high redshift rather than low

redshift, there is relatively little degeneracy between the matter density and the other

variables – but there is also much less sensitivity to the other variables. The equivalent

normalized variables to those in the top left panel of Figure 7 are Ωw/6.60,−w0/6.59, and

−wa/36.7. This insensitivity is why contours in the Ωm-Ωw or Ωm-w planes are rather

vertical for distances tied to high redshift, like baryon acoustic oscillations. Of course

since the degeneracy directions between distances tied to low redshift and those tied to

high redshift are different, these measurements are complementary. (Although not as

much in the w0-wa plane, since high redshift is essentially blind to these parameters so

tilde and regular distances have the same dependence.)

For the Hubble parameter, a nearly triple degeneracy exists between Ωm, Ωw and

wa out to z ≈ 1. Interestingly, for H there is a strong degeneracy between Ωm and w0

for z ≈ 1− 2, while the Ωw-wa degeneracy remains. It is quite difficult to observe H(z)

directly, i.e. measured relative to low redshift; more common is measurement relative

to high redshift as through baryon acoustic oscillation in the line of sight direction (see

§3.4). Here, the sensitivity to the dark energy equation of state is reduced relative to

the distance case, even apart from the degeneracies (since H does not include the z <∼ 1

lever arm).

For either d or H the correlation between energy densities is reversed from the

regular quantities, so the likelihood contour orientation in the Ωm-ΩΛ plane is reversed

as well, close to that of the CMB, another measurement tied to the high redshift universe.

The contour in Ωm-w0 is similarly reversed but not in w0-wa. Thus, all these varieties of

distance measurements have similar equation of state dependencies and, unfortunately,

no substantial orthogonality can be achieved in this plane.

To emphasize this last point, Figure 9 illustrates the degeneracy directions for

contours of constant Hubble parameter and constant matter density at various redshifts.

As we consider redshifts running from z = 0 to z ≫ 1, the isocontours only rotate

moderately. Since they always remain oriented in the same quadrant, orthogonality is

not possible.

Conversely, while not giving as much complementarity in the dark energy equation

of state as we might like, the use of different probes can be viewed as giving

complementarity in mapping the expansion history. Observations of distances out to

z = 1.7, e.g. through supernova data, plus weak lensing are basically equivalent to

mapping the expansion at z ≈ 1.5, supernovae plus CMB give it at z ≈ 0.8, while

CONTENTS 20

Figure 9. Lines of constant expansion history, i.e. Hubble parameter, are plotted

in the dark energy equation of state plane. Observations over a range of redshifts

give complementarity, but never complete orthogonality. Contours show approximate

examples of how certain combinations of distance probes effectively map different parts

of the expansion history.

supernovae plus a present matter density prior provide it at z ≈ 0.3.

In this subsection we have seen that a considerable part of the optimal approach

for mapping the expansion history is set purely by the innate cosmological dependence,

and the survey design must follow these foundations as basic science requirements. The

purpose of detailed design of successful surveys is rather to work within this framework

to minimize systematic uncertainties in the measurements, discussed in more detail in

§5. These two requirements give some of the main conclusions of this review.

3.2. Type Ia supernovae

The class of exploding stars called Type Ia supernovae (SN) become as bright as a galaxy

and can be seen to great distances. Moreover, while having a modest intrinsic scatter

in luminosity, they can be further calibrated to serve as standardized candles. Used in

this way as distance measures, studies of SN led to the discovery of the acceleration of

the universe [3, 4].

As of the beginning of 2008, over 300 SN with measurement quality suitable for

cosmological constraints had been published, representing the efforts of several survey

groups. Finding a SN is merely the first step: the time series of flux (the lightcurve),

must be measured with high signal to noise from before peak flux (maximum light) to a

month or more afterward. This must be done for multiple wavelength bands to permit

CONTENTS 21

dust and intrinsic color corrections. Additionally, spectroscopy to provide an accurate

redshift and confirmation that it is a Type Ia supernova must be obtained. See [42] for

further observational issues. (Also see [43] for the use of Type II-P supernovae.)

Once the multiwavelength fits and corrections are carried out, one derives the

distance, often spoken of in terms of the equivalent magnitude or logarithmic flux known

as the distance modulus. The distance-redshift or magnitude-redshift relation is referred

to as the Hubble diagram. This, or any other derived distance relation, can then be

compared to cosmological models.

Analysis of the world SN data sets published as of the start of 2008, comprising 307

SN, show the expansion history is consistent with a cosmological constant plus matter

universe, ΛCDM, but also with a great variety of other models [44]. Essentially no

constraints on dark energy can be placed at z > 1 and the limits on time variation

are no more stringent than the characteristic time scale being of the Hubble time or

shorter. Figure 10 shows that even for a flat universe and constant equation of state there

is considerable latitude in dark energy characteristics and that systematic uncertainties

need to be reduced for further numbers of SN to be useful.

0.0 0.1 0.2 0.3 0.4 0.5-1.5

-1.0

-0.5

0.0

Ωm

w

SNSN w/ sys

0.0 0.1 0.2 0.3 0.4 0.5-1.5

-1.0

-0.5

0.0

Ωm

w

SN + BAO + CMBSN w/ sys + BAO + CMB

Figure 10. 68.3%, 95.4%, and 99.7% confidence level contours on the matter density

and constant equation of state from the Union 2007 set of supernova distances. The left

panel shows SN only and the right panel includes current CMB and baryon acoustic

oscillation constraints. From [44].

3.3. Cosmic microwave background

The geometric distance information within the cosmic microwave background radiation

(CMB) arises from the angular scale of the acoustic peaks in the temperature power

spectrum, reflecting the sound horizon of perturbations in the tightly coupled photon-

baryon fluid at the time of recombination. If one can predict from atomic physics

CONTENTS 22

the sound horizon scale then it can be used as a standard ruler, with the angular

scale then measuring the ratio of the sound horizon s to the distance to CMB last

scattering at zlss ≈ 1089. Related, but not identical to the inverse of this distance ratio,

R =√Ωmh2 dlss is often called the reduced distance or CMB shift parameter. This gives

a good approximation to the full CMB leverage on the cosmic expansion for models near

ΛCDM [45]; for nonstandard models where the sound horizon is affected, this needs to

be corrected [44] or supplemented [46].

Because the CMB essentially delivers a single distance, it cannot effectively break

degeneracies between cosmological parameters on its own. While some leverage comes

from other aspects than the geometric shift factor (such as the integrated Sachs-Wolfe

effect), the dark energy constraints tend to be weak. However, as a high redshift distance

indicator the CMB does provide a long lever arm useful for combination with other

distance probes (effectively the z = ∞ line in Figure 9), and can strongly aid in breaking

degeneracies [47].

The reduced distance to last scattering, being measured with excellent precision –

1.8% in the case of current data and the possibility of 0.4% precision with Planck data

– defines a thin surface in cosmological parameter space (see, e.g., [48]). From Figure 11

we see that the acoustic peak structure contains substantial geometric information for

mapping the cosmological expansion. This will provide most of the information on dark

energy since cosmic variance prevents the low multipoles from providing substantial

leverage (and almost none at all to the geometric quantities, as shown by the similarity

of dashed and solid contours). Polarization information from E-modes and the cross

spectrum improves the geometric knowledge. The distance ratio dlss/s of the distance to

last scattering relative to the sound horizon is determined much better (almost 20 times

so) than the reduced distance to last scattering R =√Ωmh2 dlss, but there is substantial

covariance between these quantities (as evident in the right panel) for models near

ΛCDM (unless broken through external priors on, e.g., the Hubble constant).

The tightly constrained geometric information means that certain combinations of

cosmological parameters are well determined, but this can actually be a pitfall if one is

not careful in interpretation. Under certain circumstances, one such parameter is the

value of the dark energy equation of state at a particular redshift, w⋆ = w(z ≈ 0.4),

related to a so-called pivot redshift. If CMB data is consistent with a cosmological

constant universe, ΛCDM, then this forces w⋆ ≈ −1. However, this does not mean that

dark energy is the cosmological constant – that is merely a mirage as a wide variety

of dynamical models would be forced to give the same result. See §5.2 for further

discussion.

3.4. Baryon acoustic oscillations

The sound horizon imprinted in the density oscillations of the photon-baryon fluid,

showing up as acoustic peaks and valleys in the CMB temperature power spectrum,

also appears in the large scale spatial distribution of baryons. Since galaxies, galaxy

CONTENTS 23

148 148.5 149 149.5 150 150.5

14250

14300

14350

14400

14450

14500

s [Mpc]

d lss [

Mpc

]

96.10 96.15 96.20 96.251.67

1.68

1.69

1.70

1.71

1.72

dlss / s

R

Figure 11. Left panel: CMB data determines the geometric quantities of the distance

to last scattering dlss and the sound horizon scale s precisely, and their ratio (the

slope of the contours) – the acoustic peak angle – even more so. The contours give the

68.3% (white line) and 95.4% (black line) confidence levels for a cosmic variance limited

experiment. Outer (dark blue to light green) contours use the temperature power

spectrum, with dashed contours restricted to multipoles ℓ ≥ 40, while the inner (light

gold to dark red) contours include the E-mode polarization and TE cross-spectrum.

Right panel: Similar contours for the shift parameter R =√Ωmh2 dlss and acoustic

peak scale.

clusters, and other objects containing baryons can be observed at various redshifts,

measurements of the angular scale defined by the standard ruler of the sound horizon

provide a distance-redshift probe. Table 1 compares the acoustic oscillations in baryons

and photons.

Table 1. Comparison of signals from perturbations in the prerecombination coupled

baryon-photon fluid.

Photons Baryons

Effect CMB acoustic peaks Baryon acoustic oscillations

Scale 1 100 h−1 Mpc comoving

Base amplitude 5× 10−5 10−1

Oscillation amplitude O(1) 5%

Detection 1015/hand/sec indirect: light from < 1010 galaxies

For spatial density patterns corresponding to baryon acoustic oscillations (BAO)

transverse to the line of sight, this is an angular distance relation while radially, along

the line of sight, this is a proper distance interval, corresponding in the limit of small

interval to the inverse Hubble parameter since drprop = dz/[(1 + z)H ]. It is important

to remember that the quantities measured are actually distance ratios, i.e. d = da/s and

H = sH . These give different dependencies on the cosmological parameters than SN

distances, as rather than being ratios relative to low redshift distances they are ratios

relative to the high redshift universe (see §3.1 and Figures 7-8).

CONTENTS 24

The BAO scale was first measured with moderate statistical significance in 2005

with Sloan Digital Sky Survey data [49, 50, 51] and 2dF survey data [52]. Current

precision is 3.6% on the angular distance measurement at z = 0.35 and 6.5% at z = 0.5.

Comparing the z = 0.35 SDSS result with the 2dF result at z = 0.2 reveals some tension

[53], likely between the data sets [54] rather than from major deviation from the ΛCDM

model. Note that the quantities quoted to date are not individual reduced distances but

rather a roughly spherically averaged quantity convolving angular distance and proper

distance interval.

To obtain the proper distance interval, and hence nearly the reduced Hubble

parameter, requires accurate spectroscopic information to avoid projection of modes

along the line of sight. Photometric redshifts allow estimation of the reduced angular

distance but even that is at reduced precision until the redshift resolution approaches

spectroscopic quality [55]. Since the sound horizon scale is about 1/30 of the Hubble

radius today, assembling sufficient statistics to measure accurately the subtle signal

requires huge numbers of baryon markers over great volumes (106-109 over 1-100 Gpc3).

For early papers on the cosmological use of BAO and issues regarding the interpretation

of data, see [56, 57, 58, 59, 60, 61, 62, 63, 64, 65, 66].

3.5. Other methods

Many other probes of the expansion history and cosmic acceleration have been suggested,

in the geometry, geometry+mass, or geometry+mass+gas categories. However many

of them either lack a robust physical foundation or have yet to achieve practical

demonstration of cosmological constraints. We therefore merely list some of the

geometric methods with a few brief comments.

General tests of the expansion – These are in a certain sense the most fundamental

and interesting. One well known test is that the CMB temperature should evolve with

redshift as T ∼ 1 + z. This depends on the phase space density of the radiation being

conserved, following Liouville’s Theorem, e.g. photons do not convert into axions or

appreciably interact with other components not in equilibrium with the radiation. It

also requires the expansion to be adiabatic. Measurements to date, out to z = 3, show

agreement with T ∼ 1 + z at modest precision, and further prospects include using the

Sunyaev-Zel’dovich effect (Compton upscattering) in clusters to sample the radiation

field seen by galaxy clusters at different redshifts [67].

Another test depending on Liouville conservation is the thermodynamic, or

reciprocity, relation that can be phrased as a redshift scaling between angular distances

and luminosity distances or “third party” angular distances between points not including

the observer. Most commonly the relation is written as dl = (1 + z)2da. This has a

long history in cosmology, dating from the 1930s with Tolman [68], Ruse [69], and

Etherington [70], through the 1970s with Weinberg [71], and a general proof in terms

of the Raychaudhuri equation and the second law of thermodynamics by [72, 31]. The

thermodynamic connection is basically that if two identical blackbodies sent photons to

CONTENTS 25

each other in a cyclic process over cosmic distances then work would be done unless the

relation holds.

While the generality of these two probes as tests of the nature of the expansion

itself is intriguing, if a violation were found one might be far more likely to blame

uncertainties in the sources than a radical breakdown of the physics.

Epochs of the expansion – The expansion history can also be probed at certain

special epochs. For example, light element abundances are sensitive to the expansion

rate during primordial nucleosynthesis [73]. The CMB recombination epoch and sound

horizon are influenced by deviations from matter and radiation dominated expansion

[74, 75]. In the future, if we measure the abundance and mass of dark matter particles as

thermal relics, we may be able to constrain the expansion rate at their freezeout epoch

near perhaps 1 TeV (z = 1016) [76]. Such early constraints are particularly interesting

in multifield or multiepoch acceleration models [77].

Gamma ray bursts – As sources detectable to high redshift and insensitive to

dust extinction, some hope existed that gamma ray bursts could serve as standardized

candles to high redshift (but see §3.6 and Figure 12 regarding the limited leverage

of high redshift). However, while SN have an intrinsic flux dispersion of less than

50% that one calibrates down to 15% scatter, GRBs start with an isotropic-equivalent

energy dispersion of a factor 1000, making any calibration much more challenging.

Standardization relations were often ill defined and sensitive to environment [78] and

with better data it is now realized that much of the correlation arose from experimental

selection effects [79].

Gravitational wave sources – The intrinsic amplitude of clean systems like

inspiraling black hole binaries can be predicted from observations of the gravitational

waveform, giving a standard siren to compare to detected gravitational wave amplitude,

thus providing a distance [80, 81]. No such observations have yet been achieved.

Obtaining the redshift part of the distance-redshift relation depends on observing an

electromagnetic counterpart.

Gravitational lensing cosmography – Deflection of light by massive objects depends

on the mass distribution, the distances of source and lens from the observer, and the laws

of gravity. However, cross-correlation techniques between sources at different redshifts

have been proposed to reduce the need for knowing the mass distribution to extract

distances [82, 83, 84]. Practical demonstration has not yet occurred, but is an exciting

prospect, and gravitational lensing also serves as a growth probe (see §4.3).Environment dependent methods – Techniques that intrinsically depend on not just

a single source but its environment or group properties require greater understanding

of many astrophysical factors. Such techniques often have not yet established a clear

or robust physical motivation for the standardization essential to a distance probe.

Approaches attempted include the galaxy age-redshift relation [85, 86, 87], active

galactic nuclei radio jets [88], active galactic nuclei reverberation mapping [89, 90],

and starburst galaxies [91].

CONTENTS 26

3.6. Summary of mapping techniques

• To map the cosmological expansion accurately requires clear, robust, mature

measurement techniques.

• Observations over a range of redshifts are required to break degeneracies, or

equivalently rotate likelihood contours. Complementary probes do this as well.

• Systematics control can narrow contours and prevent biases in parameter

estimation; see §5 for further analysis and discussion.

We have seen that much of the leverage for mapping the cosmic expansion is

determined by the innate cosmology dependence on the parameters, and any survey

design must work within this framework. One implication of this is that there is a survey

depth of diminishing returns; beyond some redshift z ≈ 2 little additional leverage

accrues to measuring such high redshift distances relative to low redshift distances.

That is, although the lever arm is longer, the lever is weaker. For high redshift

distances measured relative to the high redshift universe, the lever arm is shorter, so

again improvement in constraints stalls at high redshift.

Figure 12 illustrates that once the expansion history is determined out to, say,

z ≈ 1.5 (or 2/3 the age of the universe) that it is exceedingly difficult to measure

geometrically more about dark energy with any significance. Even the simple question

of whether the equation of state has any deviation from the cosmological constant

requires percent level accuracy. Note that at high redshift the range w ∈ [−∞, 0] is

comprehensive in that w = −∞ means the dark energy density contributes not at all,

and w > 0 would mean it upsets matter domination at high redshift (limits on this are

discussed further in §4.1). Thus distance probes of the expansion history are driven

naturally by cosmology to cover the range from z = 0 to 1.5–2.

4. Growth and expansion

Within the theory of general relativity, the cosmic expansion history completely

determines the growth of matter density perturbations on large scales. On smaller

scales the state of the matter, e.g. the Jeans length defined by pressure effects, enters

as well and when density fluctuations go nonlinear then radiative and gas heating and

cooling, star formation, feedback etc. affect the formation and evolution of structure.

Here we consider only the large scale, linear regime and the role that growth observations

can play in mapping the cosmological expansion.

4.1. Growth of density perturbations

For a matter density perturbation δρ, its evolution is reduced from the Newtonian

exponential growth on the gravitational dynamical timescale tdyn = 1/√Gρ to power

law growth on a Hubble timescale by the drag induced from the cosmological expansion.

CONTENTS 27

Figure 12. Once the most recent two-thirds of the cosmological expansion is

determined, deeper measurements of distances tied to low redshift have little further

leverage. Over the physically viable range for the high redshift dark energy equation

of state w ∈ [−∞, 0], distances do not vary between models by more than 1-

2%, exceedingly difficult to measure accurately at high redshift. Cosmological

considerations alone indicate that covering the range out to z = 1.5− 2 is the optimal

strategy.

Generally,

g′′ +[

4 +1

2(lnH2)′

]

g′ +[

3 +1

2(lnH2)′ − 3

2Ωm(a)

]

g = 0, (9)

where g = (δρ/ρ)/a is the normalized growth, a prime denotes differentiation with

respect to ln a, and the dimensionless matter density Ωm(a) = Ωma−3/[H/H0]

2. Thus

the growth indeed depends only on the expansion history H (the matter density is

implicit within the high redshift, matter dominated expansion behavior).

In a matter dominated epoch, the solution is δρ/ρ ∼ a, so we defined g such

that it would be constant (which we can arrange to be unity) in such a case. In

an epoch dominated by an unclustered component with equation of state w, matter

density perturbations do not grow. Note that this is due to the small source term, not

the Hubble drag – indeed the Hubble drag term is proportional to 5 − 3w and is less

strong in a radiation dominated universe than a matter dominated universe, yet matter

density perturbations still do not grow in the radiation epoch. If one somehow fixed

Ωm(a) as one changed w, then increasing w indeed aids growth. For domination by a

component with w < 0 (not acceleration per se) there is the double whammy of reduced

source term and increased friction.

CONTENTS 28

As the universe makes the transition from being dominated by a w > 0 component

to matter domination, the solution is

δρ/ρ ∼ 1 +ΩmΩw

1

w(1 + 3w)a3w. (10)

One can see the transition from no growth δ ∼ constant to growth (with δ ∼ a

for a radiation-matter transition). For a transition from matter domination to being

dominated by a w < 0 component, to first order

δρ/ρ ∼ a

[

1− 1− w

−w(5− 6w)

ΩwΩm

a−3w

]

. (11)

Finally, for a w = 0 component that does not behave in the standard fashion, e.g.

a constant fraction F of the matter density does not clump, then

δρ/ρ ∼ a(√25−24F−1)/4. (12)

For example, a small fraction Ωe of early dark energy with w = 0 leads to δρ/ρ ∼a1−(3/5)Ωe at early times. See [92, 93, 9] for early discussion of these growth behaviors

in multicomponent universes.

Since the origin of the density perturbation source term is the Poisson equation,

basically contributing 4πGNδρ, any modification to this equation gives an effective

F (k, a) ≡ [GΩm(k, a)]/[GNΩm(a)]std that produces modified growth, where the

wavenumber k allows for spatially dependent modifications. If F is nearly constant

in time, then the growth is determined by Eq. (12). However, if the physical origin

of the modification also affects the expansion, as from a time dependent gravitational

coupling GN , then one must alter the other growth equation terms as well to obtain the

solution.

Growth measurements, being integrals from high redshift, can probe the high

redshift universe. For example, a measurement of the growth factor at, say, z = 2

that agrees with a ΛCDM model mapped out at lower redshift ensures that the high

redshift epoch of matter domination occurred substantially as expected. Specifically, if

the linear growth factor at z = 2 deviates by less than 5%, then for a monotonically

varying dark energy model described by w(a) = w0+wa(1−a), we can limit wa < 0.6 or

w(z = 2) < −0.6. We can also constrain an intermediate, transient epoch of acceleration

in the cosmic expansion. Growth measurements can place tight constraints on this

mechanism for easing the coincidence problem [94], as shown in Figure 13. To prevent

strong deviations in growth that would be evident in measurements, the period of such

acceleration must be much shorter than the characteristic Hubble timescale (and hence

apparently “unnatural”).

Overall, the growth history has the potential to constrain the expansion history.

However, there are some obstacles. The growth also depends on the initial conditions,

couplings to other components, and deviations of gravity from general relativity. For

example, in a radiation dominated universe growth should not occur, but if the

perturbation evolution has some “velocity” from a previous matter dominated epoch

it can proceed with a slow, ln a growth. (Similarly, growth persists into an accelerating

CONTENTS 29

Figure 13. Growth measurements can strongly constrain events in the cosmological

expansion such as an intermediate epoch of acceleration. The left panel show the

deviations from the ΛCDM growth, in the present total growth g0 and growth ratio

R = g(a = 0.35)/g0, for a transition to total equation of state −1 lasting a period

∆ ln a, ending at redshift zu. The right panel shows the limit on the length of the

acceleration allowed as a function of transition redshift, in order for deviations to the

total growth to be less than 10%.

epoch.) Another example is that in the prerecombination universe baryons were tightly

coupled to photons, preventing growth [95]. Finally, it is quite difficult to detect the

matter perturbations per se, instead we observe light from galaxies – a biased tracer of

the density field. Separating the absolute growth factor from the bias, or the relative

growth from an evolution in bias, is nontrivial [96]. Weak gravitational lensing (see §4.3below) offers a way around these issues.

Measurements of growth are presently not clean probes of the expansion history,

especially since, as Eq. (9) indicates, precision knowledge of Ωm is required to truly

use this approach to map H(z). Redshift distortion measurements of the growth of the

matter velocity field (rather than density), employing d ln g/d ln a, have similar issues.

See [97, 98] for good overviews of the current status of growth measurements.

4.2. Abundance tests

Under gravitational instability, mass aggregates and galaxies and clusters of galaxies

form, evolving in number and mass. The abundance as a function of redshift of particular

classes of objects will involve (among other variables) the initial power spectrum,

astrophysical processes such as dissipation and feedback, and the expansion rate of

the universe, as well as a robust observational proxy for mass. Two famous examples of

using galaxy abundances as cosmological probes are in a 1975 paper “An Accelerating

CONTENTS 30

Universe” [99] and in 1986 “Measurement of the Mass Density of the Universe” [100]. As

the titles suggest, the data led to opposite conclusions and each was quickly recognized

to be showing more about astrophysical evolution of the objects than the cosmological

expansion history.

New generations of experiments are underway to count sources as a function of

redshift and mass, selecting samples through their detection in the Sunyaev-Zel’dovich

effects, X-ray flux, optical flux, or weak gravitational lensing shear. Currently, data is

insufficient to give robust cosmological constraints on the expansion history. Note that

observations do not actually measure an abundance per comoving volume, dN(m)/dV ,

which would show how structures grow in mass while their (nonevolving) numbers should

stay constant. Rather they see dN/dz which involves distances through an extra factor

of dV/dz. Thus observed abundances do not give a pure growth probe but mix geometry

and growth – both an advantage and disadvantage.

4.3. Gravitational lensing

The gravitational potentials of massive structures deflect, focus, and shear bundles

of light rays from more distant sources. The particular shear pattern combines

information from both the lensing mass and the “focal length” defined by the source

and lens distances from the observer. To date, most such data provides leverage on

a combination of the matter density Ωm and the rms amplitude of mass fluctuations

σ8. A few surveys so far quote further cosmological constraints – on the dark energy

parameters in combination with other data sets [101], or on the presence of growth

[102, 103, 104, 105, 106].

A number of interesting concepts for using gravitational lensing to map the

cosmological expansion exist, though practicality has yet to be demonstrated. Both the

magnification or convergence fields and the shear field carry cosmological information.

For use as a geometric probe, one must separate astrophysical or instrumental effects,

deconvolve the lens gravitational potential model, and measure the redshifts or redshift

distribution of the lenses and sources accurately. See [82, 83, 84] for discussion of weak

lensing as a geometric probe. Similarly, lensing offers promise to probe the growth

history. Lensing of the CMB, where the source redshift is known, is another interesting

application [107]. See [108] for far more discussion of weak lensing and its many uses

than given here.

4.4. Testing gravity

4.4.1. First steps Although the focus of this article is on mapping the cosmological

expansion, since the growth history depends on both the expansion and the laws of

gravity one could consider using expansion plus growth measurements to test the

gravitational framework. This is an exciting possibility that has recently received

considerable attention in the literature (see, e.g., [109, 37, 110, 111] for early work),

although current data cannot provide significant constraints.

CONTENTS 31

In addition to comparing predictions of specific theories of gravity to data, one basic

approach is parametrization of gravitational cosmological effects. While one could define

cosmological parameters Ωgrm, w

gr0 , w

gra , say, derived from fitting the growth history and

contrast these with those derived from pure expansion measurements, one could also

keep the physics of cosmic expansion expressed through the well-defined (effective)

parameters Ωm, w0, wa and look for physical effects from gravity beyond Einstein

relativity in separate, clearly interpretable gravitational parameters.

The gravitational growth index γ was designed specifically to preserve this

distinction of physical phenomena [112]. Here the matter density linear growth factor

g = (δρ/ρ)/a is written

g(a) = e∫

a

0(da′/a′) [Ωm(a′)γ−1]. (13)

See [112, 113, 114] for discussion of the accuracy, robustness, and basis of this form.

However, gravity does more than affect the growth: it affects the light deflection

law in lensing and the relation between the matter density and velocity fields. Since

two potentials enter generically, in the time-time and space-space terms of the metric,

[115, 116, 117, 118, 119, 120, 121, 122] and others have proposed two or more parameters

for testing extensions to general relativity (GR). These may involve the ratio of metric

potentials (unity for GR), as a cosmological generalization of the parametrized post-

Newtonian quantity, and their difference (zero for GR), related to the anisotropic stress.

4.4.2. Problems parametrizing beyond-Einstein gravity The two gravitational

parameter approach is an exciting idea that deserves active exploration. Here, however,

to balance the literature we undertake a more critical overview of this program.

Difficulties to overcome include:

(i) Reduction of two functions to two (or a small number) of parameters;

(ii) Spatial dependence, e.g. strong coupling or Vainshtein scales, or quantities living

in real space vs. Fourier space;

(iii) Covariance among deviations in gravitational coupling, metric potentials, and fluid

anisotropic stress.

We give very brief assessments of each item. For the first point, a successful proof

of concept is the gravitational growth index, which for a wide class of theories adds a

single (constant) parameter to describing linear growth. [114] explains why this works

under certain conditions, basically requiring small deviations from standard cosmological

expansion and gravity. Many theories indeed have small deviations, but these are often

so small they become problematic to detect; other theories with larger deviations are

either ruled out by, e.g., solar system tests or have scale dependence requiring multiple

parameters.

Regarding the second point, in the parametrized post-Newtonian (PPN) formalism

the ratio of metric potentials is defined in real space, e.g. Φ(r)/Ψ(r). This is not

equivalent to defining a parameter η = Φk/Ψk in Fourier space. Given some function

CONTENTS 32

f(r) = Φ(r)+Ψ(r), it is not generically true that a nonlinear function F (f(r)) depends

in the same manner on Φ and Ψ. Thus many power spectra, e.g. involving 〈(Φ + Ψ)2〉for the integrated Sachs-Wolfe effect, will not follow the simple formalism.

Spatial variation also does not preserve the functional dependence; for example,

while the light deflection angle depends on the specific combination Φ(r) + Ψ(r), the

actual deflection at impact parameter b is not a function only of Φ(b) + Ψ(b). Consider

lensing in DGP [123] gravity, where

Φ(r) =m

r+

nφr3

(14)

Ψ(r) =m

r+

nψr3

(15)

where the n’s are related to extradimensional quantities. (Many IR modifications of

gravity have this form, but note it is very different from the Yukawa form so general

parametrization of spatial dependence may be difficult). Thus f(r) takes the form

2m/r + n/r3 and some function, say,

F (f(r0)) =∫ ∞

r0

dr

rf(r) =

2m

r0+

n

3r306= C [Φ(r0) + Ψ(r0)], (16)

for any constant C. More concretely, for deflection of light at impact parameter b from

a point mass (extending [124]), we find the deflection angle

αb = 2 [Φ + Ψ]b − 6 [Ψ− Φ]b, (17)

breaking the functional form Φ+Ψ with an extra term looking like a spurious anisotropic

stress. Thus observations of lensing deflection do not tell us about a simple parameter

η in the form Φ (1 + η), as we might hope.

Another issue concerns interdependence among different aspects of gravitational

modifications. Variation in gravitational coupling may well occur alongside anisotropic

stress, as in scalar-tensor theories. In Poisson’s equation, therefore, one cannot define a

Fourier space mass power spectrum 〈δ2k〉 because the physical source involves 〈(Geffδk)2〉

and it may not be clear how the interaction of these two quantities affects the result.

This is reminiscent of varying constant theories where one can calculate the effects

of a varying fine structure constant, say, but does not have a unified framework for

understanding how the electron-proton mass ratio, say, varies simultaneously, and

hence affects observations as well [125]. Confusion also arises between gravitational

modifications and fluid microphysics, e.g. anisotropic stress of a component, coupling

between components, or finite sound speed (see, e.g., [126]).

4.4.3. Levels of discovery In summary, there is a rich array of challenges for a

program seeking to parameterize beyond Einstein gravity. One ray of hope is that

different observations depend differently on the quantities, some of which are outlined

in Table 2 (also see [120]). In the theory realm, SN serve as pure geometric probes of

the cosmological expansion, immune to the uncertainties imposed by these additional

parameters (even G(t), see e.g. [131]), providing the most unambiguous understanding.

CONTENTS 33

Table 2. Theory systematics occur even for some geometric probes, the cleanest

astrophysically. For example, light deflection depends on the metric potentials as Φ+Ψ,

plus separating out the mass model, and baryon acoustic oscillations are sensitive to

the validity of the usual CDM growth, hence affected by Φ and Ψ (see, for example,

[127] for difficulties in the braneworld case), new density perturbations due to the

sound speed cs (see, e.g., [128]) and anisotropic stress πs, variation of the gravitational

coupling G with scale and time, and altered fluctuations due to matter coupling Γ (see,

e.g., [129, 130]).

Probe Theory Systematic (dominant) Theory Systematic (potential)

SN Ia — —

WL Φ+Ψ cs, πs, G(k, t)

BAO Φ, Ψ, cs, Γ πs, G(k, t)

This point is worth considering: all other methods in use rely on understanding the rest

of the dark sector – dark matter and gravity – as they seek to explore dark energy. As

well though, we should keep in mind the insight by Richard Feynman:

“Yesterday’s sensation is today’s calibration, and tomorrow’s background.”

Mapping the cosmological expansion with a simple, robust, and geometric method

like supernovae provides a firm foundation as we also must probe deeper, with techniques

that have more complex – richer – dependence on further variables revealing the

microphysics and testing gravity. Mutual support among methods will be key to yielding

understanding.

5. Systematics in data and theory

Accurate mapping of the cosmological expansion requires challenging observations over

a broad redshift range with precision measurements. Technological developments such

as large format CCDs, large telescopes, and new wavelength windows make such surveys

possible. In addition to obtaining large data sets of measurements, we must also address

systematic uncertainties in measurements and astrophysical source properties. Beyond

that, for accurate mapping we must consider systematics in the theoretical interpretation

and the data analysis. Here we present some brief illustrations of the impact of such

systematics on mapping the expansion.

5.1. Parameterizing dark energy

In extracting cosmological parameters from the data, one wants parameters that are

physically revealing, that can be fit precisely, and that accurately portray the key

physical properties. For exploration of many different parametrizations and approaches

see [18] and references therein. Any functional form for the expansion, e.g. the

dark energy equation of state w(z), runs the risk of limiting or biasing the physical

interpretation, so one can also consider approaches such as binned equations of state,

CONTENTS 34

defined as tophats in redshift, say, or decomposition into orthogonal basis functions

or principal component analysis (see, e.g., [132]). However, for a finite number of

components or modes this does not solve all the problems, and introduces some new ones

such as model dependence and uncertain signal-to-noise criteria (see [133] for detailed

discussion of these methods).

Indeed, even next generation data restricts the number of well-fit parameters for

the dark energy equation of state to two [134], greatly reducing the flexibility of basis

function expansion or bins. Here, we concentrate on a few comments regarding two

parameter functions and how to extract clear, robust physics from them.

For understanding physics, two key properties related to the expansion history and

the nature of acceleration are the value of the dark energy equation of state and its time

variation w′ = dw/d lna. These can be viewed as analogous to the power spectral tilt

and running of inflation. The two parameter form

w(a) = w0 + wa(1− a), (18)

based on a physical foundation by [41], has been shown to be robust, precise, bounded,

and widely applicable, able to match accurately a great variety of dark energy physics.

See [135] for tests of its limits of physical validity.

One of the main virtues of this form is its model independence, serving as a global

form able to cover reasonably large areas of dark energy phase space, in particular both

classes of physics discussed in §2.3 – thawing and freezing behavior. We can imagine,

however, that future data will allow us to zero in on a particular region of phase space,

i.e. area of the w-w′ plane, as reflecting the physical origin of acceleration. In this case,

we would examine more restricted, local parametrizations in an effort to distinguish

more finely between physics models.

First consider thawing models. One well-motivated example is a pseudoscalar field

[136], a pseudo-Nambu Goldstone boson (PNGB), which can be well approximated by

w′ = F (1 + w) (19)

w(a) = − 1 + (1 + w0) aF , (20)

where F is inversely proportional to the PNGB symmetry breaking energy scale f .

Scalar fields, however, thawing in a matter dominated universe, must at early times

evolve along the phase space track w′ = 3(1 + w) [17], where w departs from −1 by a

term proportional to a3. One model tying this early required behavior to late times and

building on the field space parametrization of [137] is the algebraic model of [18],

1 + w = (1 + w0) ap

(

1 + b

1 + ba−3

)1−p/3

, (21)

with parameters w0, p (b = 0.3 is fixed). Figure 14 illustrates these different behaviors

and shows matching by the global (w0, wa) model, an excellent approximation to w(z)

out to z >∼ 1. More importantly, it reproduces distances to all redshifts to better than

0.05%. The key point, demonstrated in [18], is that use of any particular one of these

parametrizations does not bias the main physics conclusions when testing consistency

CONTENTS 35

with the cosmological constant or the presence of dynamics. Thus one can avoid a

parametrization-induced systematic.

Figure 14. Within the thawing class of physics models one can hope to distinguish

specific physics origins. A canonical scalar field evolves along the early matter

dominated universe behavior shown by the “matter limit” curve, then deviates as

dark energy begins to dominate. Such thawing behavior is well fit by the algebraic

model. A pseudoscalar field (PNGB) evolves differently. Either can be moderately

well fit by the phenomenological (w0, wa) parametrization for the recent universe.

For freezing models, we can consider the extreme of the early dark energy model

of [138], with 3% contribution to the energy density at recombination, near the upper

limit allowed by current data [139]. This model is specifically designed to represent

dark energy that scales as matter at early times, transitioning to a strongly negative

equation of state at later times. If future data localizes the dark energy properties to

the freezing region of phase space, one could compare physical origins from this model

(due to dilaton fields) with, say, the Hα phenomenological model of [140] inspired by

extra dimensions. Again a key point is that the global (w0, wa) parametrization does

not bias the physical conclusions, matching even the specialized, extreme early dark

energy model to better than 0.02% in distance out to z ≈ 2.

Parametrizations to be cautious about are those that arbitrarily assume a fixed

high redshift behavior, often setting w = −1 above some redshift. These can strongly

bias the physics [133, 135]. As discussed in §2, interesting and important physics clues

may reside in the early expansion behavior.

Bias can also ensue by assuming a particular functional form for the distance or

Hubble parameter [141]. Even when a form is not assumed a priori, differentiating

CONTENTS 36

imperfect data (e.g. to derive the equation of state) leads to instabilities in the

reconstruction [142, 143]. To get around this, one can attempt to smooth the data,

but this returns to the problems of assuming a particular form and in fact can remove

crucial physical information. While the expansion history is innately smooth (also see

§6.2), extraction of cosmological parameters involves differences between histories, which

can have sharper features.

5.2. Mirage of Lambda

As mentioned in §3.3, interpreting the data without fully accounting for the possibility of

dynamics can bias the theoretical interpretation. We highlight here the phenomenon of

the “mirage of Λ”, where data incapable of precisely measuring the time variation of the

equation of state, i.e. with the quality expected in the next five years, can nevertheless

apparently indicate with high precision that w = −1.

Suppose the distance to CMB last scattering – an integral measure of the equation

of state – matches that of a ΛCDM model. Then under a range of circumstances, low

redshift (z <∼ 1) measurements of distances can appear to show that the equation of state

is within 5% of the cosmological constant value, even in the presence of time variation.

Figure 15 illustrates the sequence of physical relations leading to this mirage.

Matching the distance to last scattering imposes a relation between the cosmological

parameters, here shown as a family of (w0, wa) models holding other parameters fixed.

The convergence in their distances beginning at a ≈ 0.65 is associated with a similar

convergence in the fractional dark energy density, and a matching in w(z) at a ≈ 0.7.

Note that even models with substantial time variation, wa ≈ 1, are forced to have

w(z ≈ 0.4) = −1, i.e. look like the cosmological constant. This is dictated by the innate

cosmological dependences of distance and is robust to different parameter choices; see

[135] for more details. (Note the matching in ΩDE(a) forced at a ≈ 0.5 has implications

for the linear growth factor and nonlinear power spectrum, as explored in [144].)

To see beyond the mirage of Λ, or test its reality, requires measurements capable

of directly probing the time variation with significant sensitivity (hence extending out

to z ≈ 1.7 as shown in §3.1). Current and near term experiments that may show

w ≈ −1 to a few percent can induce a false sense of security in Λ. In particular, the

situation is exacerbated by the pivot or decorrelation redshift (where the equation of

state is measured most precisely) of such experiments probing to z ≈ 1 being close to

the matching redshift imposed on w(z); so, given CMB data consistent with Λ, such

experiments will measure w = −1 with great precision, but possibly quite inaccurately.

See Figure 16 for a simulation of the possible data interpretation in terms of the mirage

of Λ from an experiment capable of achieving 1% minimum uncertainty on w (i.e.

w(zpivot)). Clearly, the time variation wa is the important physical quantity allowing

us to see through the mirage, and the key science requirement for next generation

experiments.

CONTENTS 37

Figure 15. Matching the distance to CMB last scattering between dark energy models

leads to convergence and crossover behaviors in other cosmological quantities. The top

left panel illustrates the convergence in the distance-redshift relation, relative to the

ΛCDM case, for models with w0 ranging from −0.8 to −1.2 and corresponding time

variation wa. The top right panel illustrates the related convergence and crossover in

the dark energy density ΩDE(a), and the bottom panel shows how the CMB matching

necessarily leads to a crossover with w = −1 at the key redshift for sensitivity of low

redshift experiments. This crossover in w(z) leads to the mirage of Λ, and is impelled

by the physics not the functional form.

5.3. Inhomogeneous data sets

Turning from theory to data analysis, another source of systematics that can lead

to improper cosmological conclusions are heterogeneous data sets. This even holds

with data all for a single cosmological probe, e.g. distances. Piecing together distances

CONTENTS 38

Figure 16. If CMB data are consistent with ΛCDM, this can create a mirage of Λ for

lower redshift distance data even if the dark energy has substantial time variation. The

curves show simulated 68% confidence regions for w(z) for two different CMB-matched

models. The value of the equation of state w(z) = −1 necessarily, for each one at a

redshift close to the “sweet spot” or pivot redshift. Experiments insufficiently precise

to see time variation will think w = −1 to high precision (the width of the narrow

waist at z ≈ 0.38, here 1%) even if the true behavior is drastically different.

measured with different instruments under different conditions or from different source

samples opens the possibilities of miscalibrations, or offsets, between the data.

While certainly an issue when combining, say, supernova distances with baryon

acoustic oscillation distances, or gravitational wave siren distances, we illustrate

the implications even for heterogeneous supernova samples. An offset between the

magnitude calibrations can have drastic effects on cosmological estimation (see, e.g.,

[145]). For example, very low redshift (z < 0.1) supernovae are generally observed

with very different telescopes and surveys than higher redshift (z > 0.1) ones. Since

the distances in the necessary high redshift (z ≈ 1 − 1.7) sample require near infrared

observations from space then the crosscalibration with the local sample (which requires

very wide fields and more rapid exposures) requires care. The situation is exacerbated

if the space sample does not extend down to near z ≈ 0.1 − 0.3 and a third data set

intervenes. This gives a second crosscalibration needed to high accuracy.

Figure 17 demonstrates the impact on cosmology, with the magnitude offset leading

to bias in parameter estimation. If there is only a local distance set and a homogeneous

space set extending from low to high redshift, with crosscalibration at the 0.01 mag

level, then the biases take on the values at the left axis of the plot: a small fraction

CONTENTS 39

of the statistical dispersion. However, with an intermediate data set, the additional

heterogeneity from matching at some further redshift zmatch (with the systematic taken

to be at the 0.02 mag level to match a ground based, non-spectroscopic experiment to

a space based spectroscopic experiment) runs the risk of bias in at least one parameter

by of order 1σ. Thus cosmological accuracy advocates as homogeneous a data set as

possible, ideally from a single survey.

Figure 17. Heterogeneous datasets open issues of imperfect crosscalibration, modeled

here as magnitude offsets ∆m. One scenario involves calibration between a local

(z < 0.1) spectroscopic set and a uniform survey extending from z ≈ 0.1 − 1.7. This

imposes cosmological parameter biases given by the intersection of the curves with the

left axis. Another scenario takes the high redshift data to consist of two, heterogeneous

sets with an additional offset ∆m at some intermediate matching redshift zmatch.

(When zmatch = 0.1 this corresponds to the first scenario with no extra offset.)

Similar heterogeneity and bias can occur in baryon acoustic oscillation surveys

mapping the expansion when the selection function of galaxies varies with redshift. If

the power spectrum shifts between samples, due for example to different galaxy-matter

bias factors between types of galaxies or over redshift, then calibration offsets in the

acoustic scale lead to biases in the cosmological parameters. Again, innate cosmology

informs survey design, quantitatively determining that a homogeneous data set over the

redshift range is advantageous.

5.4. Miscalibrated standard

Miscalibration involving the basic standard, i.e. candle luminosity or ruler scale, has

a pernicious effect biasing the expansion history mapping. This time we illustrate the

CONTENTS 40

point with baryon acoustic oscillations. If the sound horizon s is improperly calibrated,

with an offset δs (for example through early dark energy effects in the prerecombination

epoch [146, 147]), then every baryon acoustic oscillation scale measurement d(z) =

d(z)/s and H(z) = sH(z) will be miscalibrated. Due to the redshift dependence of the

untilded quantities, the offset will vary with redshift, looking like an evolution that can

be confused with a cosmological model biased from the reality.

To avoid this pitfall, analysis must include a calibration parameter for the sound

horizon (since CMB data does not uniquely determine it [148, 147]), in exact analogy to

the absolute luminosity calibration parameter M required for supernovae. That is, the

standard ruler must be standardized; assuming standard CDM prerecombination for

the early expansion history blinds the analysis to the risk of biased cosmology results.

The necessary presence of a standard ruler calibration parameter, call it S, leads toan increase in the w0-wa contour area, and equivalent decrease in the “figure of merit”

defined by that area, by a factor 2.3. Since we do not know a priori whether the high

redshift universe is conventional CDM (e.g. negligible early dark energy or coupling),

neglecting S for BAO is as improper as neglecting M for supernova standard candle

calibration. (Without the need to fit for the low redshift calibration M, SN would enjoy

an improvement in “figure of merit” by a factor 1.9, similar to the 2.3 that BAO is given

when neglecting the high redshift calibration S.)For supernovae, in addition to the fundamental calibration of the absolute

luminosity, experiments must tightly constrain any evolution in the luminosity [149].

This requires broadband flux data from soon after explosion to well into the decline

phase, and spectral data over a wide frequency range. Variations in supernova

properties that do not affect the corrected peak magnitude do not affect the cosmological

determination.

5.5. Malmquist bias

Distance measurements from the cosmological inverse square law of flux must avoid

threshold effects where only the brightest sources at a given distance can be detected,

known as Malmquist bias. Suppose the most distant, and hence dimmest, sources were

close to the detection threshold. We can treat this selection effect as a shift in the mean

magnitude with a toy model,

∆m = −Az − z⋆

0.95− z⋆, z > z⋆ . (22)

This then propagates into a bias on the cosmological parameter fit to the data.

Consider a data set of some 1000 supernovae from z = 0−1, with the Malmquist bias

setting in at z⋆ = 0.8 (where ground based spectroscopy begins to get quite time intensive

and many spectral features move into the near infrared). The bias in a cosmological

parameter relative to its uncertainty is then

δp

σp= (1.5, 1.0, 1.2)× A

0.1 mag(23)

CONTENTS 41

for p = (Ωm, w0, wa). Thus, the Malmquist bias must be limited to less than 0.05 mag

at z = 0.95 to prevent significant bias in the derived cosmological model. In fact, this

is not a problem for a well designed supernova survey since the requirement of mapping

out the premaximum phase of the supernova ensures sensitivity to fluxes at least two

magnitudes below detection threshold.

5.6. Other issues

In addition to the theory interpretation and data analysis systematics discussed in this

section, recall the fundamental theory systematics of §4.4.3 and Table 2. We finish with

a very brief mention of some other selected data and data analysis systematics issues of

importance that are often underappreciated and that must be kept in mind for proper

survey design and analysis.

• Sample variance: Along a given line of sight, the local distance measures anchoring

the Hubble diagram can be influenced by coherent velocity flows, throwing off the

derived cosmology [150, 151, 152]. The local distances should therefore be well

into the Hubble flow and the sources distributed widely on the sky. In addition,

the mass distribution along the line of sight may not be representative of the

homogeneous model and gravitational lensing can lead to coherent magnification

effects (relevant for standardized candles) and alterations of the measured three

dimensional clustering (important for baryon acoustic oscillations). See, e.g.,

[153, 154]. For these reasons and others, “pencil beam” surveys can be fraught

with systematics and are poor survey design.

• Analytic marginalization: The calibration parameter, e.g. M combining the

absolute luminosity and Hubble constant in the case of supernovae, is often referred

to as a nuisance parameter but its proper treatment is essential. Although in some

χ2 formulas for the distance-redshift relation it is not written explicitly, it is implicit

and cannot be ignored. More subtle is the issue of analytic marginalization over it –

this must be used with great care as the distribution of M is actually non-Gaussian

due to interaction with other supernovae peak magnitude fitting quantities (such

as the lightcurve width and color terms) [44]. Further subtleties exist between

marginalization and minimization in a multidimensional fit space [155, 152], and

most analysis from raw data to quoted parameters actually employs minimization

techniques.

• Extinction priors: Since the dimming and reddening due to dust effects on

supernovae are one-sided (i.e. dust does not increase the flux), they are highly

non-Gaussian and must be treated with care. Any deviation between an assumed

prior for extinction and the truth, that is not constant in redshift, can bias the

cosmology results. See Figure 18 and the handy systematics calculator SMock

[156] for examples. Several analysis techniques avoid this pitfall by fitting for dust

and intrinsic color globally, without assuming a prior, though this requires high

quality data over several wavelength bands.

CONTENTS 42

-2.5

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

2.5

-2 -1.8 -1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0w0

w′=

wa

⁄ 2

Prior with AV bias

Ellipticals only

No prior on extinction

Prior with AV+RV bias

39% CL contours (1σ projected)σ(Ωm) = 0.04 prior

×

Figure 18. Assuming a prior on dust properties (AV and RV ) in order to reduce

extinction errors can cause systematic deviations in the magnitudes. These give strong

biases to the equation of state, leading to a false impression of a transition from w < −1

to w > −1. To avoid this systematic, one can use samples with minimal extinction

(elliptical galaxy hosted supernovae) or obtain precise multiwavelength data that allows

for fitting the dust and color properties. Based on [157].

6. Future prospects

The main uncertainty in our future ability to map the cosmological expansion is the

level of control of systematics we can achieve. In a reductio ad absurdum we can say

that since one supernova explodes every second in the universe we might measure 107

per year, giving distance accuracies of 10%/√107 = 0.003% in ten redshift slices over

a ten year survey; or counting every acoustic mode in the universe per redshift slice

spanning the acoustic scale λ determines the power spectrum to λ3/(H−2λ) = (1/30)2

or 0.1%; or measuring weak lensing shear across the full sky with 0.1” resolution takes

individual 1% shears to the 0.0003% level. Statistics is not the issue: understanding of

systematics is.

6.1. Data and systematics

However it is much easier to predict the future of statistical measurements than

systematic uncertainties. It is difficult to estimate what the future prospects really

are. Large surveys are being planned assuming that uncertainties will be solved – or

if the data cannot be used for accurately mapping the cosmic expansion it will still

CONTENTS 43

prove a cornucopia to many fields of astrophysics. Moreover, an abundance of surveys

mentioned in the literature are various levels of vaporware: many have never passed a

national peer review or been awarded substantial development funding or had their costs

reliably determined. Rather than our listing such possibilities, the reader can peruse

national panel reports such as the ESA Cosmic Vision program [158], US National

Academy of Sciences’ Beyond Einstein Program Assessment Committee report [159], or

the upcoming US Decadal Survey of Astronomy and Astrophysics [160].

The current situation regarding treatment of systematics is mixed. In supernova

cosmology, rigorous identification and analysis of systematics and their effects on

parameter estimation is (almost) standard procedure. In weak lensing there exists the

Shear Testing Programme [161], producing and analyzing community data challenges.

Other techniques have less organized systematics analysis, where it exists, and the crucial

rigorous comparison of independent data sets (as in [44]) is rare. Control methods such

as blind analysis are also rare. We cannot yet say where the reality will lie between

the unbridled statistical optimism alluded to in the opening of this section and current,

quite modest measurements. Future prospects may be bright but considerable effort is

still required to realize them.

6.2. Mapping resolution

Since prediction of systematics control is difficult, let us turn instead to intrinsic limits

on our ability to map the expansion history – limits that are innate to cosmological

observables. The expansion history a(t), like distances, is an integral over H = d ln a/dt,

and does not respond instantaneously to the evolution of energy density (which in turn

is an integral over the equation of state). As seen in Figures 7-8, the cosmological kernel

for the observables is broad – no fine toothed comb exists for studying the expansion.

This holds even where models have rapid time variation w′ > 1, as in the e-fold transition

model, or when considering principal components (see [162, 133] for illustrations).

(This resolution limit on mapping the expansion is not unique to distances – the

growth factor is also an integral measure and the broad kernel of techniques like weak

lensing is well known. The Hubble parameter determined through BAO, say, requires

a redshift shell thickness ∆z >∼ 0.2 to obtain sufficient wave modes for good precision,

limiting the mapping resolution.)

Thus, due to the innate cosmological dependences of observables, plus additional

effects such as Nyquist and statistics limits of wavemodes in a redshift interval

and coherence of systematics over redshift [162], we cannot expect mapping of the

cosmological expansion with finer resolution than ∆z ≈ 0.2 from next generation data.

6.3. Limits on cosmic doomsday

Finally, we turn from the expansion history to the expansion future. As pointed out in

§2.5, to determine the fate of our universe we must not only map the past expansion

CONTENTS 44

but understand the nature of the acceleration before we can know the universe’s destiny

– eternal acceleration, fading of dark energy, or recollapse.

We do not yet have that understanding, but suppose we assume that the linear

potential model of dark energy [38], perhaps the simplest alternative to a cosmological

constant, is correct. Then we can estimate the time remaining before the fate of

recollapse: the limit on cosmic doomsday. The linear potential model effectively has

a single equation of state parameter and is well approximated by the family with

wa = −1.5(1 + w0) when w0 is not too far from −1. Curves of the expansion history

(and future) are illustrated in the left panel of Figure 19; the doomsday time from the

present is given by

tdoom ≈ 0.5H−10 (1 + w0)

−0.8. (24)

The current best constraints from data (cf. [163] for future limits) appear in the

right panel of Figure 19, corresponding to tdoom > 24 Gy at 68% confidence level. It is

obviously of interest to us to know whether the universe will collapse and how long until

it does, so accurate determination of wa is important! The difference between doomsday

in only two Hubble times from now and in three (a whole extra Hubble time!) is only

a difference of 0.12 in wa.

-1 1 2 3t

2

4

6

8

a

0.0 0.1 0.2 0.3 0.4 0.50.0

0.5

1.0

1.5

2.0

2.5

Ωm

H0*

t

SN w/ sys + BAO + CMBBAOCMBSN w/ sys

Figure 19. The future expansion history in the linear potential model has a recollapse,

or cosmic doomsday. Left panel: Precision measurements of the past history are

required to distinguish the future fate, with the five curves corresponding to five

values for the potential slope (the uppermost curve is for a flat potential, i.e. a

cosmological constant). From [163]. Right panel: Current data constraints estimate

cosmic doomsday will occur no sooner than ∼1.5 Hubble times from now. From [164].

CONTENTS 45

7. Conclusions

Mapping the cosmological expansion is a key endeavour in our quest for understanding

physics – gravitation and other forces, spacetime, the vacuum – and the origin, evolution,

and future of the universe. The Big Bang model of a hot, dense, early universe expanding

and adiabatically cooling, and forming structure through gravitational instability from

primordial seed perturbations, is remarkably successful and simple. The concordance

cosmology is (close to) spatially flat and 13.7 billion years old. This clear statement

represents a substantial advance of our knowledge over a decade ago.

The discovery of the acceleration of the cosmic expansion created a renaissance

of cosmological exploration, offering the hope of connecting quantum physics and

gravitation, extra dimensions and the nature of spacetime, while severing the bonds

between geometry and destiny. This opens up completely two premier questions in

science: the origin and the fate of the universe. To understand the nature of the

gravitationally repulsive dark energy pulling the universe apart, we must map the

cosmological expansion in greater detail and accuracy than ever envisioned.

We have shown that a considerable part of the optimal approach for mapping the

expansion history is set purely by the innate cosmological dependences. Observational

programs must follow these foundations as basic science requirements: in particular the

need for a wide range of redshifts, z ≈ 0− 2 and robust anchoring to either low or high

redshift. Our capabilities for measuring the expansion through direct geometric probes

are increasing, and techniques are continuing to develop. There are no short cuts –

detailed design of successful surveys works within this framework with the purpose to

minimize systematic uncertainties in the measurements.

Every technique has systematic uncertainties, appearing in multiple guises. While

observational systematics are most familiar, arising even in purely geometric techniques,

equally important are issues in the data analysis, e.g. combining heterogeneous data

sets, and in the theoretical interpretation, e.g. susceptibility to biases from assumptions

of high redshift behavior or of non-expansion physics (perturbation growth behavior,

coupling, etc.).

We have given several concrete examples of the effects of systematic uncertainties

for various probes. These provide a cautionary tale for survey design, as we are now

entered into the systematics dominated data era – and may well soon approach the

theory systematics era. We cannot rely on assumptions that any part of the dark sector

is simple and ignorable while we concentrate on another aspect. For true progress, we

emphasized the role of complementarity in building from robust, clean answers to more

complex investigations. Probes employing growth of structure give windows on both

expansion per se and gravitational laws, and we commented that the excitement of

testing general relativity is equally matched with the challenge of creating a framework

for analysis.

Acceleration of the cosmic expansion heralds a revolution in physics, if we can

characterize and understand it. To comprehend this new aspect of the universe, we

CONTENTS 46

must map the expansion not only at recent epochs, but encompass the early universe.

Once we have garnered sufficient understanding, the prize is answering the question of

the fate of the universe, now unbound from the question of the cosmic geometry. The

good news is that we likely have at least 24 billion years to do so!

Acknowledgments

I thank the Aspen Center for Physics for a superb working environment, and Los Alamos

National Laboratory and the Santa Fe 07 workshop, University of Heidelberg, and the

Dark Cosmology Centre and Niels Bohr Summer Institute, for hospitality. I gratefully

acknowledge contributions by Georg Robbers, Ramon Miquel, and David Rubin of

Figures 11, 18, and 19 respectively. Colleagues too numerous to mention helped form

my thinking on this wide ranging topic, but I will single out Bob Wagoner for laying the

foundations. This work has been supported in part by the Director, Office of Science,

Department of Energy under grant DE-AC02-05CH11231.

References

[1] Focus issue on dark energy 2008, Gen. Rel. Grav. 40, xx-xx

[2] Bernstein J & Feinberg G (eds) 1986, Cosmological Constants: Papers in Modern Cosmology

(Columbia U Press: New York)

[3] Riess A G et al. 1998, Astron. J. 116, 1009-1038

[4] Perlmutter S et al. 1999, ApJ 517, 565-586

[5] Frieman J A, Turner M S, Huterer D 2008, Ann. Rev. Astron. Astroph. 46, in press

[6] Sandage A 1962, ApJ 136, 319-333

[7] McVittie G 1962, ApJ 136, 334-338

[8] Linder E V 1991, at http://supernova.lbl.gov/˜evlinder/drift.html

[9] Linder E V 1997, First Principles of Cosmology (London: Addison-Wesley)

[10] Uzan J-P, Bernardeau F, Mellier Y 2007, arXiv:0711.1950

[11] Blandford R D, Amin M, Baltz E A, Mandel K, Marshall P J 2005, ASP Conf. Ser. 339, 27-38

[arXiv:astro-ph/0408279]

[12] Rapetti D, Allen S W, Amin M A, Blandford R D 2007, MNRAS 375, 1510-1520

[13] Shapiro C & Turner M S 2006, ApJ 649, 563-569

[14] Weinberg S 1989, Rev. Mod. Phys. 61, 1-23

[15] Carroll S M 2001, Liv. Rev. Rel. 4, 1

[16] Bousso R 2008, Gen. Rel. Grav. 40, xxx

[17] Caldwell R R & Linder E V 2005, Phys. Rev. Lett. 95, 141301

[18] Linder E V 2008, Gen. Rel. Grav. 40, xxx

[19] Olive K A & Peacock J A 2007, at http://pdg.lbl.gov/2007/reviews/bigbangrpp.pdf

[20] Lahav O & Liddle A R 2007, at http://pdg.lbl.gov/2007/reviews/hubblerpp.pdf

[21] Sachs R K 1961, Proc. Roy. Soc. London A 264, 309-338

[22] Kristian J & Sachs R K 1966, ApJ 143, 379-399

[23] Gunn J E 1967, ApJ 150, 737-753

[24] Jacobs M W, Linder E V, Wagoner R V 1992, Phys. Rev. D 45, R3292-R3295

[25] Stoeger W R, Araujo M E, Gebbie T 1997, ApJ 476, 435-439 ; erratum ApJ 522, 559

[26] Feynman R P 1964, unpublished lecture notes

[27] Zel’dovich Y B 1964, Sov. Ast. 8, 13-16

[28] Dyer C C & Roeder R C 1972, ApJ 174, L115-L117

CONTENTS 47

[29] Linder E V 1998, arXiv:astro-ph/9801122

[30] Ishibashi A & Wald R M 2006, Class. Quant. Grav. 23, 235-250

[31] Linder E V 1988, Astron. Astroph. 206, 190-198

[32] Lineweaver C H & Davis T M 2005, Sci. Am. 292, 36-45

[33] Krauss L M & Starkman G D 2000, ApJ 531, 22-30

[34] Krauss L M & Scherrer R J 2007, Gen. Rel. Grav. 39, 1545-1550

[35] Rindler W 2006, Relativity: Special, General, and Cosmological (Oxford: Oxford U. Press)

[36] Caldwell R R, Kamionkowski M, Weinberg N N 2003, Phys. Rev. Lett. 91, 071301

[37] Linder E V 2004, Phys. Rev. D 70, 023511

[38] Linde A 1987, in Three Hundred Years of Gravitation, ed. S W Hawking & W Israel (Cambridge:

Cambridge U. Press), p. 604-630

[39] Steinhardt P J & Turok N 2002, Science 296, 1436-1439

[40] Tegmark M, Eisenstein D J, Hu W, Kron R 1998, arXiv:astro-ph/9805117

[41] Linder E V 2003, Phys. Rev. Lett. 90, 091301

[42] Kim A G 2008, in Dark Energy, ed. P Ruiz-Lapuente (Cambridge: Cambridge U. Press), in press

[43] Nugent P et al. 2006, ApJ 645, 841-850

[44] Kowalski M et al. 2008, ApJ submitted

[45] Efstathiou G & Bond J R 1999, MNRAS 304, 75-97

[46] Wright E L 2007, ApJ 664, 633-639

[47] Frieman J A, Huterer D, Linder E V, Turner M S 2003, Phys. Rev. D 67, 083505

[48] Hu W, Huterer D, Smith K M 2006, ApJ 650, L13-L16

[49] Eisenstein D J et al. 2005, ApJ 633, 560-574

[50] Hutsi G 2006, Astron. Astroph. 449, 891-902

[51] Padmanabhan, N et al. 2007, MNRAS 378, 852-872

[52] Cole S et al. 2005, MNRAS 362, 505-534

[53] Percival W J et al. 2007, MNRAS 381, 1053-1066

[54] Sanchez A G & Cole S 2007, arXiv:0708.1517

[55] Padmanabhan N 2008, in preparation

[56] Meiksin A, White M, Peacock J A 1999, MNRAS 304, 851-864

[57] Cooray A 2002, in A New Era in Cosmology, ASP Conf. Proc. 283, 162-163

[arXiv:astro-ph/0112418]

[58] Eisenstein D 2002, ASP Conf. Proc. 280, 35-42

[59] Blake C & Glazebrook K 2003, ApJ 594, 665-673

[60] Linder E V 2003, Phys. Rev. D 68, 083504

[61] Hu W & Haiman Z 2003, Phys. Rev. D 68, 063004

[62] Seo H-J & Eisenstein D J 2003, ApJ 598, 720-740

[63] Glazebrook K & Blake C 2005, ApJ 631, 1-20

[64] White M 2005, Astropart. Phys. 24, 334-344

[65] Linder E V 2005, arXiv:astro-ph/0507308

[66] Seo H-J & Eisenstein D J 2005, ApJ 633, 575-588

[67] Battistelli E S et al. 2002, ApJ 580, L101-L104

[68] Tolman R C 1930, Proc. Nat. Acad. Sci. 16, 511-520

[69] Ruse H S 1932, Proc. Roy. Soc. Edinburgh 52, 183-194

[70] Etherington I M H 1933, Phil. Mag. 15, 761

[71] Weinberg S 1972, Gravitation and Cosmology (Wiley: New York)

[72] Ellis G F R 1971, in General Relativity and Cosmology, ed. R Sachs (London: Academic Press),

p. 104-182

[73] Carroll S M & Kaplinghat M 2002, Phys. Rev. D 65, 063507

[74] Zahn O & Zaldarriaga M 2003, Phys. Rev. D 67, 063002

[75] Chan K C & Chu M-C 2007, Phys. Rev. D 75, 083521

[76] Chung, D J H, Everett L L, Kong K, Konstantin T 2007, JHEP 10, 016

CONTENTS 48

[77] Griest K 2002, Phys. Rev. D 66, 123501

[78] Friedman A S & Bloom J S 2005, ApJ 627, 1-25

[79] Butler N R, Kocevski D, Bloom J S, Curtis J L 2007, ApJ 671, 656-677

[80] Schutz B F 1986, Nature 323, 310-311

[81] Holz D E & Hughes S A 2005, ApJ 629, 15-22

[82] Jain B & Taylor A 2003, Phys. Rev. Lett. 91, 141302

[83] Bernstein G & Jain B 2004, ApJ 600, 17-25

[84] Zhang J, Hui L, Stebbins A 2005, ApJ 635, 806-820

[85] Jimenez R, Verde L, Treu T, Stern D 2003, ApJ 593, 622-629

[86] Simon J, Verde L, Jimenez R 2005, Phys. Rev. D 71, 123001

[87] Verkhodanov O V, Parijskij Y N, Starobinsky A A 2005, Bull. SAO 58, 5-15 [arXiv:0705.2776]

[88] Daly R et al. 2007, arXiv:0710.5112

[89] Kaspi S 2007, ASP Conf. Ser. 373, 13-22 [arXiv:0705.1722]

[90] Teerikorpi P 2005, arXiv:astro-ph/0510382

[91] Siegel E R et al. 2005, MNRAS 356, 1117-1122

[92] Fry J N 1985, Phys. Lett. B 158, 211-214

[93] Linder E V 1988, MPA internal research note

[94] Linder E V 2006, Astropart. Phys. 26, 16-21

[95] Meszaros P 1974, Astron. Astroph. 37, 225-228

[96] Bernardeau F, Colombi S, Gaztanaga E, Scoccimarro R 2002, Phys. Rep. 367, 1-248

[97] Tegmark M et al. 2006, Phys. Rev. D 74, 123507

[98] Mcdonald P et al. 2006, ApJS 163, 80-109

[99] Gunn, J E & Tinsley, B M 1975, Nature 257, 454-457

[100] Loh, E D & Spillar E J 1986, ApJ 307, L1-L4

[101] Jarvis M, Jain B, Bernstein G, Dolney D 2006, ApJ 644, 71-79

[102] Benjamin J et al. 2007, MNRAS 381, 702-712

[103] Semboloni E et al. 2006, Astron. Astroph. 452, 51-61

[104] Massey R et al. 2007, ApJS 172, 239-253

[105] Fu L et al. 2007, arXiv:0712.0884

[106] Dore et al. 2007, arXiv:0712.1599

[107] Hu W, Holz D E, Vale C 2007, Phys. Rev. D 76, 127301

[108] Hoekstra H & Jain B 2008, Ann. Rev. Nucl. Part. Sci. 58, in press

[109] Lue A, Scoccimarro R, Starkman G 2004, Phys. Rev. D 69, 124015

[110] Knox L, Song Y-S, Tyson J A 2006, Phys. Rev. D 74, 023512

[111] Ishak M, Upadhye A, Spergel D 2006, Phys. Rev. D 74, 043513

[112] Linder E V 2005, Phys. Rev. D 72, 043529

[113] Huterer D & Linder E V 2007, Phys. Rev. D 75, 023519

[114] Linder E V & Cahn R N 2007, Astropart. Phys. 28, 481-488

[115] Caldwell R, Cooray A, Melchiorri A 2007, Phys. Rev. D 76, 023507

[116] Zhang P, Liguori M, Bean R, Dodelson S 2007, Phys. Rev. Lett. 99, 141302

[117] Amendola L, Kunz M, Sapone D 2007, arXiv:0704.2421

[118] Hu W & Sawicki I 2007, Phys. Rev. D 76, 104043

[119] Amin M, Wagoner R, Blandford R 2007, arXiv:0708.1793

[120] Jain B & Zhang P 2007, arXiv:0709.2375

[121] Bertschinger E & Zukin P 2008, arXiv:0801.2431

[122] Daniel S F, Caldwell R R, Cooray A, Melchiorri A 2008, arXiv:0802.1068

[123] Dvali G, Gabadadze G, Porrati M 2000, Phys. Lett. B 485, 208-214

[124] Keeton C R & Petters A O 2006, Phys. Rev. D 73, 104032

[125] Uzan, J-P 2003, Rev. Mod. Phys. 75, 403-455

[126] Kunz M & Sapone D 2007, Phys. Rev. Lett. 98, 121301

[127] Song Y-S, Sawicki I, Hu W 2007, Phys. Rev. D 75, 064003

CONTENTS 49

[128] DeDeo S, Caldwell R R, Steinhardt P J 2003, Phys. Rev. D 67, 103509

[129] Amendola L 2004, Phys. Rev. D 69, 103524

[130] Mangano G, Melchiorri A, Serra P, Cooray A, Kamionkowski M 2006, Phys. Rev. D 74, 043517

[131] Linder E V 2003, at http://supernova.lbl.gov/˜evlinder/gdotsn.pdf

[132] Huterer D & Starkman G 2003, Phys. Rev. Lett. 90, 031301

[133] de Putter R & Linder E V 2007, arXiv:0710.0373

[134] Linder E V & Huterer D 2005, Phys. Rev. D 72, 043509

[135] Linder E V 2007, arXiv:0708.0024

[136] Frieman J A, Hill C T, Stebbins A, Waga I 1995, Phys. Rev. Lett. 75, 2077-2080

[137] Crittenden R, Majerotto E, Piazza F 2007, Phys. Rev. Lett. 98, 251301

[138] Doran M & Robbers G 2006, JCAP 0606, 026

[139] Doran M, Robbers G, Wetterich C 2007, Phys. Rev. D 75, 023003

[140] Dvali G & Turner M S 2003, arXiv:astro-ph/0301510

[141] Jonsson J, Goobar A, Amanullah R, Bergstrom L 2004, JCAP 0409, 007

[142] Tegmark M 2002, Phys. Rev. D 66, 103507

[143] Linder E V 2004, Phys. Rev. D 70, 061302

[144] Francis M J, Lewis G F, Linder E V 2007, MNRAS 380, 1079-1086

[145] Linder E V 2006, Astropart. Phys. 26, 102-110

[146] Doran M, Stern S, Thommes E 2007, JCAP 0704, 015

[147] Linder E V & Robbers G, in draft

[148] Eisenstein D J & White M 2004, Phys. Rev. D 70, 103523

[149] Branch D, Perlmutter S, Baron E, Nugent P 2001, arXiv:astro-ph/0109070

[150] Hui L & Greene P B 2006, Phys. Rev. D 73, 123526

[151] Cooray A & Caldwell R R 2006, Phys. Rev. D 73, 103002

[152] Neill J D, Hudson M J, Conley A 2007, ApJ 661, L123-L126

[153] Cooray A, Huterer D, Holz D E 2007, Phys. Rev. Lett. 96, 021301

[154] LoVerde M, Hui L, Gaztanaga E 2007, arXiv:0708.0031

[155] Upadhye A, Ishak M, Steinhardt P J 2005, Phys. Rev. D 72, 063501

[156] Nordin J, Goobar A, Jonsson J 2008, JCAP in press [arXiv:0801.2484] ;

further information and calculator at http://www.physto.se/˜nordin/smock

[157] Linder E V & Miquel R 2004, Phys. Rev. D 70, 123516

[158] Cosmic Vision 2007, at http:/sci.esa.int/science-e/www/area/index.cfm?fareaid=100

[159] BEPAC 2007, at http://books.nap.edu/catalog.php?record id=12006

[160] Astro2010 Survey, at http://www7.nationalacademies.org/bpa/Astro2010.html, in preparation

[161] Shear Testing Programme, at http://www.physics.ubc.ca/˜heymans/step/cosmic shear test.html

Heymans C et al. 2006, MNRAS 368, 1323-1339 ;

Massey R et al. 2007, MNRAS 376, 13-38

[162] Linder E V 2007, Phys. Rev. D 75, 063502

[163] Kallosh R, Kratochvil J, Linde A, Linder E V, Shmakova M 2003, JCAP 0310, 015

[164] Rubin D, Kowalski M, Linder E V, Perlmutter S 2008, in draft